[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (131)

Search Parameters:
Keywords = memantine

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 1910 KiB  
Article
Development and Validation of a Simple UV–HPLC Method to Quantify the Memantine Drug Used in Alzheimer’s Treatment
by Débora Nunes, Tânia G. Tavares, Frenacisco Xavier Malcata, Joana A. Loureiro and Maria Carmo Pereira
Pharmaceuticals 2024, 17(9), 1162; https://doi.org/10.3390/ph17091162 - 2 Sep 2024
Viewed by 351
Abstract
Memantine, a non-competitive NMDA receptor antagonist, is used to treat Alzheimer’s disease. Therefore, loading memantine in nanoparticles (NPs) could be an essential tool to improve the treatment effectiveness while reducing drug toxicity. Even though some approaches have been described to quantify memantine, none [...] Read more.
Memantine, a non-competitive NMDA receptor antagonist, is used to treat Alzheimer’s disease. Therefore, loading memantine in nanoparticles (NPs) could be an essential tool to improve the treatment effectiveness while reducing drug toxicity. Even though some approaches have been described to quantify memantine, none reported optimized methods using high-performance liquid chromatography resorting to ultraviolet detection (UV–HPLC) to determine encapsulation in NPs. The present research developed a HPLC method using pre-column derivatization for quantitatively analyzing memantine hydrochloride in NPs. Memantine was derivatized using 9-fluorenylmethyl chloroformate (FMOC). The developed method was fully validated regarding suitability, specificity, linearity, sensitivity, precision, accuracy, and robustness according to the International Conference on Harmonisation of Technical Requirements for Registration of Pharmaceuticals for Human Use guidelines. The retention time of memantine was 11.393 ± 0.003 min, with a mean recovery of 92.9 ± 3.7%. The new chromatographic method was validated and found to respond linearly over 5–140 μg/mL, with a high coefficient of determination. Intraday precision lay between 3.6% and 4.6%, and interday precision between 4.2% and 9.3%. The stability of memantine was also tested at 4 °C and −20 °C, and no signs of decay were found for up to 6 months. The new method was properly validated and proved simple, sensitive, specific, accurate, and precise for determining memantine encapsulation efficiency in lipid NPs. Greenness was evaluated, presenting a final score of 0.45. In the future, this methodology could also be applied to quantify memantine in different nanoformulations. Full article
(This article belongs to the Special Issue Therapeutic Drug Monitoring and Adverse Drug Reactions)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of memantine hydrochloride.</p>
Full article ">Figure 2
<p>Optimization of the derivatization reaction condition in terms of (<b>A</b>) molar ratio of FMOC: memantine, (<b>B</b>) pH of borate buffer, (<b>C</b>) reaction temperature, and (<b>D</b>) reaction time. Results represent the mean ± SD, n = 3.</p>
Full article ">Figure 3
<p>Chromatograms of (<b>A</b>) blank solution (FMOC in acetonitrile) and (<b>B</b>) derivatized memantine solution.</p>
Full article ">Figure 4
<p>Calibration curve of derivatized memantine. Results represent the mean ± SD, n = 3.</p>
Full article ">Figure 5
<p>Stability analysis of memantine at 4 °C and −20 °C for 6 months. Results represent the mean ± SD, n = 3.</p>
Full article ">Figure 6
<p>Reaction scheme of the derivatization of memantine with FMOC.</p>
Full article ">Figure 7
<p>Pictogram of the HPLC method in Greenness study using the Analytical GREEnness calculator.</p>
Full article ">
15 pages, 2894 KiB  
Article
Memantine and the Kynurenine Pathway in the Brain: Selective Targeting of Kynurenic Acid in the Rat Cerebral Cortex
by Renata Kloc and Ewa M. Urbanska
Cells 2024, 13(17), 1424; https://doi.org/10.3390/cells13171424 - 26 Aug 2024
Viewed by 508
Abstract
Cytoprotective and neurotoxic kynurenines formed along the kynurenine pathway (KP) were identified as possible therapeutic targets in various neuropsychiatric conditions. Memantine, an adamantane derivative modulating dopamine-, noradrenaline-, serotonin-, and glutamate-mediated neurotransmission is currently considered for therapy in dementia, psychiatric disorders, migraines, or ischemia. [...] Read more.
Cytoprotective and neurotoxic kynurenines formed along the kynurenine pathway (KP) were identified as possible therapeutic targets in various neuropsychiatric conditions. Memantine, an adamantane derivative modulating dopamine-, noradrenaline-, serotonin-, and glutamate-mediated neurotransmission is currently considered for therapy in dementia, psychiatric disorders, migraines, or ischemia. Previous studies have revealed that memantine potently stimulates the synthesis of neuroprotective kynurenic acid (KYNA) in vitro via a protein kinase A-dependent mechanism. Here, the effects of acute and prolonged administration of memantine on brain kynurenines and the functional changes in the cerebral KP were assessed in rats using chromatographic and enzymatic methods. Five-day but not single treatment with memantine selectively activated the cortical KP towards neuroprotective KYNA. KYNA increases were accompanied by a moderate decrease in cortical tryptophan (TRP) and L-kynurenine (L-KYN) concentrations without changes in 3-hydroxykynurenine (3-HK) levels. Enzymatic studies revealed that the activity of cortical KYNA biosynthetic enzymes ex vivo was stimulated after prolonged administration of memantine. As memantine does not directly stimulate the activity of KATs’ proteins, the higher activity of KATs most probably results from the increased expression of the respective genes. Noteworthy, the concentrations of KYNA, 3-HK, TRP, and L-KYN in the striatum, hippocampus, and cerebellum were not affected. Selective cortical increase in KYNA seems to represent one of the mechanisms underlying the clinical efficacy of memantine. It is tempting to hypothesize that a combination of memantine and drugs could strongly boost cortical KYNA and provide a more effective option for treating cortical pathologies at early stages. Further studies should evaluate this issue in experimental animal models and under clinical scenarios. Full article
Show Figures

Figure 1

Figure 1
<p>The effects of an acute (<b>A</b>) and 5-day treatment (<b>B</b>) with memantine on the levels of selected kynurenines in the cerebral cortex. TRP—tryptophan, KYNA—kynurenic acid, L-KYN—L-kynurenine, 3-HK—3-hydroxykynurenine. Data are shown as the percentage of mean control values set at 100%. Statistical significance * <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. control (receiving saline injection).</p>
Full article ">Figure 2
<p>The effects of acute (<b>A</b>) and a 5-day treatment (<b>B</b>) with memantine on the levels of selected kynurenines in the hippocampus of rats. Data are shown as the percentage of mean control values set as 100%. TRP—tryptophan, KYNA—kynurenic acid, L-KYN—L-kynurenine, 3-HK—3-hydroxykynurenine.</p>
Full article ">Figure 3
<p>The effects of acute (<b>A</b>) and a 5-day treatment (<b>B</b>) with memantine on the levels of selected kynurenines in the striatum of rats. Data are shown as the percentage of mean control values set as 100%. TRP—tryptophan, KYNA—kynurenic acid, L-KYN—L-kynurenine, 3-HK—3-hydroxykynurenine.</p>
Full article ">Figure 4
<p>The effects of acute (<b>A</b>) and a 5-day treatment (<b>B</b>) with memantine on the levels of selected kynurenines in the cerebellum of rats. Data are shown as the percentage of mean control values set as 100%. TRP—tryptophan, KYNA—kynurenic acid, L-KYN—L-kynurenine, 3-HK—3-hydroxykynurenine.</p>
Full article ">Figure 5
<p>The effects of acute (<b>A</b>) or a 5-day (<b>B</b>) administration of memantine on the ex vivo activity of kynurenine aminotransferases (KATs) I and II. Data are shown as the percentage of mean control values set at 100%. Statistical significance **** <span class="html-italic">p</span> &lt; 0.001 vs. control.</p>
Full article ">
10 pages, 190 KiB  
Case Report
Sustained Cognitive Improvement in Alzheimer’s Disease Patients Following a Precision Medicine Protocol: Case Series
by Dale E. Bredesen, Mary Kay Ross and Stephen Ross
Biomedicines 2024, 12(8), 1776; https://doi.org/10.3390/biomedicines12081776 - 6 Aug 2024
Viewed by 766
Abstract
Arguably, the most important parameter in treating cognitive decline associated with Alzheimer’s disease is the length of time in which improvement, if achieved at all, is sustained. However, monotherapies such as donepezil and memantine are associated with a more rapid decline than no [...] Read more.
Arguably, the most important parameter in treating cognitive decline associated with Alzheimer’s disease is the length of time in which improvement, if achieved at all, is sustained. However, monotherapies such as donepezil and memantine are associated with a more rapid decline than no treatment in patients over multi-year follow-ups. Furthermore, anti-amyloid antibody treatment, which at best simply slows decline, is associated with accelerated cerebral atrophy, resulting in earlier dementia-associated brain volumes for those treated at the MCI stage than untreated patients. In contrast, a precision medicine approach, in which the multiple potential drivers of cognitive decline are identified for each patient and then targeted with a personalized protocol (such as ReCODE), has led to documented improvements in patients with cognitive decline, but long-term follow-up (>5 years) has not been reported previously. Therefore, here, we report sustained cognitive improvement, in some cases for over a decade, in patients treated with a precision medicine protocol—something that has not been reported in patients treated with anti-cholinesterase, glutamate receptor inhibitory, anti-amyloid, or other therapeutic methods. These case studies warrant long-term cohort studies to determine how frequently such sustained cognitive improvements occur in patients treated with precision medicine protocols. Full article
(This article belongs to the Section Neurobiology and Clinical Neuroscience)
9 pages, 252 KiB  
Review
All GLP-1 Agonists Should, Theoretically, Cure Alzheimer’s Dementia but Dulaglutide Might Be More Effective Than the Others
by Jeffrey Fessel
J. Clin. Med. 2024, 13(13), 3729; https://doi.org/10.3390/jcm13133729 - 26 Jun 2024
Viewed by 1856
Abstract
Addressing the dysfunctions of all brain cell types in Alzheimer’s disease (AD) should cure the dementia, an objective that might be achieved by GLP-1 agonist drugs, because receptors for GLP-1 are present in all of the main brain cell types, i.e., neurons, oligodendroglia, [...] Read more.
Addressing the dysfunctions of all brain cell types in Alzheimer’s disease (AD) should cure the dementia, an objective that might be achieved by GLP-1 agonist drugs, because receptors for GLP-1 are present in all of the main brain cell types, i.e., neurons, oligodendroglia, astroglia, microglia, endothelial cells and pericytes. This article describes the benefits provided to all of those brain cell types by GLP-1 agonist drugs. The article uses studies in humans, not rodents, to describe the effect of GLP-1 agonists upon cognition, because rodents’ brains differ from those of humans in so many ways that results from rodent studies may not be totally transferable to humans. Commercially available GLP-1 agonists have mostly shown either positive effects upon cognition or no effects. One important reason for no effects is a reduced rate of entering brain parenchyma. Dulaglutide has the greatest entry to brain, at 61.8%, among the available GLP-1 agonists, and seems to offer the best likelihood for cure of AD. Although there is only one study of cognition that used dulaglutide, it was randomized, placebo controlled, and very large; it involved 8828 participants and showed significant benefit to cognition. A clinical trial to test the hypothesis that dulaglutide may cure AD should have, as its primary outcome, a 30% greater cure rate of AD by dulaglutide than that achieved by an equipoise arm of, e.g., lithium plus memantine. Full article
(This article belongs to the Special Issue Potential Cures of Alzheimer's Dementia)
17 pages, 3460 KiB  
Article
γ-Oryzanol from Rice Bran Antagonizes Glutamate-Induced Excitotoxicity in an In Vitro Model of Differentiated HT-22 Cells
by Li-Chai Chen, Mei-Chou Lai, Tang-Yao Hong and I-Min Liu
Nutrients 2024, 16(8), 1237; https://doi.org/10.3390/nu16081237 - 21 Apr 2024
Viewed by 1351
Abstract
The excessive activation of glutamate in the brain is a factor in the development of vascular dementia. γ-Oryzanol is a natural compound that has been shown to enhance brain function, but more research is needed to determine its potential as a treatment for [...] Read more.
The excessive activation of glutamate in the brain is a factor in the development of vascular dementia. γ-Oryzanol is a natural compound that has been shown to enhance brain function, but more research is needed to determine its potential as a treatment for vascular dementia. This study investigated if γ-oryzanol can delay or improve glutamate neurotoxicity in an in vitro model of differentiated HT-22 cells and explored its neuroprotective mechanisms. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h then given γ-oryzanol at appropriate concentrations or memantine (10 µmol/L) for another 24 h. Glutamate produced reactive oxygen species and depleted glutathione in the cells, which reduced their viability. Mitochondrial dysfunction was also observed, including the inhibition of mitochondrial respiratory chain complex I activity, the collapse of mitochondrial transmembrane potential, and the reduction of intracellular ATP levels in the HT-22 cells. Calcium influx triggered by glutamate subsequently activated type II calcium/calmodulin-dependent protein kinase (CaMKII) in the HT-22 cells. The activation of CaMKII-ASK1-JNK MAP kinase cascade, decreased Bcl-2/Bax ratio, and increased Apaf-1-dependent caspase-9 activation were also observed due to glutamate induction, which were associated with increased DNA fragmentation. These events were attenuated when the cells were treated with γ-oryzanol (0.4 mmol/L) or the N-methyl-D-aspartate receptor antagonist memantine. The results suggest that γ-oryzanol has potent neuroprotective properties against glutamate excitotoxicity in differentiated HT-22 cells. Therefore, γ-oryzanol could be a promising candidate for the development of therapies for glutamate excitotoxicity-associated neurodegenerative diseases, including vascular dementia. Full article
(This article belongs to the Special Issue The Role of Micronutrients in Neurodegenerative Disease)
Show Figures

Figure 1

Figure 1
<p>Differentiation transformed into a mature neuron and expressed the NMDA receptor in HT-22 cells. (<b>A</b>) HT-22 cell morphology observed before differentiation (left panel) and after 48 h in a differentiation medium (right panel). After 48 h of culture in differentiation medium, HT-22 cells developed a neuron-like morphology, characterized by the presence of extended neurites. Black arrows indicate neurites. Original magnification: 200×; scale bar: 50 μm. (<b>B</b>) Differentiation conditions induce the expression of the NMDA receptor in HT-22 cells after 48 h. NMDA receptor subunit 1 (NMDAR1) was detected as the NMDA receptor. β-actin was used as loading control. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 compared to the undifferentiated group.</p>
Full article ">Figure 2
<p>γ-Oryzanol alleviates glutamate-induced neuronal damage. (<b>A</b>) Cell morphology, (<b>B</b>) cell viability, (<b>C</b>) neurite outgrowth, and (<b>D</b>) protein levels and phosphorylation degree of NMDA receptor were evaluated when the differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (Ory) at 0.2 (Ory 0.2), 0.3 (Ory 0.3), and 0.4 (Ory 0.4) mmol/L or memantine (10 µmol/L) for another 24 h. The cell morphology was examined using an Zeiss Axiovert 135 inverted phase-contrast microscope at ×100 magnification. Cell viability was determined with a CCK-8 assay and expressed as a percentage of untreated cells taken as a control group. The total number of neurites per cell was determined by counting the number of neurites directly from the soma using Image J software (version 1.6.0). Protein levels and phosphorylation degree of NMDAR1 are shown via representative immunoblots. β-actin was used as loading control. The ratio between phosphoprotein and total protein of NMDAR1 (p-NMDAR1 and NMDAR1) was calculated. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 3
<p>γ-Oryzanol protects differentiated HT-22 cells against glutamate-induced oxidative stress. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 mmol/L) or memantine (10 µmol/L) for another 24 h. (<b>A</b>) The ROS fluorescence intensity is expressed as a percentage of the untreated control cells. (<b>B</b>) The activities of SOD, GSH-Px, CAT, and GSH content were normalized to the corresponding protein concentration for each group. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 4
<p>γ-Oryzanol suppressed glutamate-induced intracellular calcium overload and CaMKII activation. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 µmol/L) or memantine (10 µmol/L) for another 24 h. (<b>A</b>) The fluorescence intensities for intracellular Ca<sup>2+</sup> was measured via live cell imaging. Original magnification: 200×; scale bar: 50 μm. Represents bar diagram of relative fluorescence intensity. (<b>B</b>) Representative immunoblots depicting the protein levels and phosphorylation degree of CaMKII, β-actin was used as loading control. The ratio of p-CaMKII to total CaMKII was calculated. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 5
<p>γ-Oryzanol prevents glutamate-induced mitochondrial dysfunction. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 µmol/L) or memantine (10 µmol/L) for another 24 h. (<b>A</b>) The mitochondrial transmembrane potential has been measured with the JC-1 fluorescence probe. (<b>B</b>) The bioluminescent detection of ADP and ATP levels was used to measure the ADP/ATP ratio in cells with a commercial assay kit. (<b>C</b>) The oxidation of NADH was measured to determine the activity of mitochondrial complex I. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 6
<p>γ-Oryzanol downregulates ASK1/JNK signaling pathway induced by glutamate. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 µmol/L) or memantine (10 µmol/L) for another 24 h. Protein expression and extent of phosphorylation on (<b>A</b>) ASK1, (<b>B</b>) MKK4, (<b>C</b>) MKK7, (<b>D</b>) JNK, (<b>E</b>) c-Jun, and (<b>F</b>) c-Fos were analyzed by Western blot. The degree of phosphorylation was calculated as a ratio of the total protein. β-actin was used as loading control. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 7
<p>γ-Oryzanol declined cytochrome c-initiated caspase cascade activation induced by glutamate. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 µmol/L) or memantine (10 µmol/L) for another 24 h. (<b>A</b>) The concentrations of cytochrome c in both mitochondrial and cytosolic fractions were determined using an immunoassay. (<b>B</b>) Protein expression of Apaf-1, procaspase-9, cleaved caspase-9, procaspase-3, cleaved caspase-3, proPARP, and cleaved PARP were analyzed by Western blot. Β-actin was used as loading control. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">Figure 8
<p>γ-Oryzanol reversed Bcl-2 protein downregulation and DNA fragmentation induced by glutamate. The differentiated HT-22 cells were treated with 0.1 mmol/L glutamate for 24 h and received γ-oryzanol (0.4 µmol/L) or memantine (10 µmol/L) for another 24 h. (<b>A</b>) The expression of Bcl-2 and Bax protein were measured by Western blot. β-actin was used as loading control. The ratio of relative intensities in Bcl-2 to Bax (Bcl-2/Bax) was reported. (<b>B</b>) The extent of apoptosis was quantified using the ELISA kit to detect DNA fragments associated with cytoplasmic histones. The results are shown as the mean ± SD of five independent experiments (<span class="html-italic">n</span> = 5) performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the untreated control group (control). <sup>c</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.01 compared to the data from glutamate-stimulated cells that received vehicle treatment.</p>
Full article ">
14 pages, 2140 KiB  
Review
Assessing the Impact of Agents with Antiviral Activities on Transmembrane Ionic Currents: Exploring Possible Unintended Actions
by Geng-Bai Lin, Chia-Lung Shih, Rasa Liutkevičienė, Vita Rovite, Edmund Cheung So, Chao-Liang Wu and Sheng-Nan Wu
Biophysica 2024, 4(2), 128-141; https://doi.org/10.3390/biophysica4020009 - 27 Mar 2024
Cited by 1 | Viewed by 928
Abstract
As the need for effective antiviral treatment intensifies, such as with the coronavirus disease 19 (COVID-19) infection, it is crucial to understand that while the mechanisms of action of these drugs or compounds seem apparent, they might also interact with unexplored targets, such [...] Read more.
As the need for effective antiviral treatment intensifies, such as with the coronavirus disease 19 (COVID-19) infection, it is crucial to understand that while the mechanisms of action of these drugs or compounds seem apparent, they might also interact with unexplored targets, such as cell membrane ion channels in diverse cell types. In this review paper, we demonstrate that many different drugs or compounds, in addition to their known interference with viral infections, may also directly influence various types of ionic currents on the surface membrane of the host cell. These agents include artemisinin, cannabidiol, memantine, mitoxantrone, molnupiravir, remdesivir, SM-102, and sorafenib. If achievable at low concentrations, these regulatory effects on ion channels are highly likely to synergize with the identified initial mechanisms of viral replication interference. Additionally, the immediate regulatory impact of these agents on the ion-channel function may potentially result in unintended adverse effects, including changes in cardiac electrical activity and the prolongation of the QTc interval. Therefore, it is essential for patients receiving these related agents to exercise additional caution to prevent unnecessary complications. Full article
Show Figures

Figure 1

Figure 1
<p>Graph demonstrating that upon exposure to the compounds or drugs described in this paper, their interactions with these ionic currents (e.g., <span class="html-italic">I</span><sub>Na</sub>, <span class="html-italic">I</span><sub>h</sub>, <span class="html-italic">I</span><sub>K(erg)</sub>, <span class="html-italic">I<sub>K</sub></span><sub>(M)</sub>, <span class="html-italic">I<sub>K</sub></span><sub>(S)</sub>, <span class="html-italic">I<sub>K</sub></span><sub>(IR)</sub>, <span class="html-italic">I<sub>K</sub></span><sub>(DR)</sub>, and <span class="html-italic">I</span><sub>MEP</sub>) will, in turn, interfere with the growth and replication of the virus. * indicates that <span class="html-italic">I</span><sub>MEP</sub> is stimulated by the presence of remdesivir (RDV).</p>
Full article ">Figure 2
<p>Impact of agents with antiviral activities on <span class="html-italic">I</span><sub>K(erg)</sub> or <span class="html-italic">I</span><sub>K(S)</sub> inhibition, and their influence on ventricular action potential ((<b>A</b>), red) fand electrocardiogram ((<b>B</b>), black). The arrowhead on the left side of (<b>A</b>) indicates the 0 mV level. The horizontal and vertical black bars in the bottom right corner of (<b>A</b>,<b>B</b>) represent time and voltage scales, respectively. The red dashed curve in (<b>A</b>,<b>B</b>) illustrates the potential effects (denoted using horizontal blue open arrowheads) of these agents, specifically the prolongation of ventricular action potential (<b>A</b>) and QTc interval (<b>B</b>), respectively.</p>
Full article ">
18 pages, 14629 KiB  
Article
NMDA Receptor Antagonist Memantine Ameliorates Experimental Autoimmune Encephalomyelitis in Aged Rats
by Biljana Bufan, Ivana Ćuruvija, Veljko Blagojević, Jelica Grujić-Milanović, Ivana Prijić, Tatjana Radosavljević, Janko Samardžić, Milica Radosavljevic, Radmila Janković and Jasmina Djuretić
Biomedicines 2024, 12(4), 717; https://doi.org/10.3390/biomedicines12040717 - 23 Mar 2024
Viewed by 1593
Abstract
Aging is closely related to the main aspects of multiple sclerosis (MS). The average age of the MS population is increasing and the number of elderly MS patients is expected to increase. In addition to neurons, N-methyl-D-aspartate receptors (NMDARs) are also expressed [...] Read more.
Aging is closely related to the main aspects of multiple sclerosis (MS). The average age of the MS population is increasing and the number of elderly MS patients is expected to increase. In addition to neurons, N-methyl-D-aspartate receptors (NMDARs) are also expressed on non-neuronal cells, such as immune cells. The aim of this study was to investigate the role of NMDARs in experimental autoimmune encephalomyelitis (EAE) in young and aged rats. Memantine, a non-competitive NMDAR antagonist, was administered to young and aged Dark Agouti rats from day 7 after immunization. Antagonizing NMDARs had a more favourable effect on clinical disease, reactivation, and apoptosis of CD4+ T cells in the target organ of aged EAE rats. The expression of the fractalkine receptor CX3CR1 was increased in memantine-treated rats, but to a greater extent in aged rats. Additionally, memantine increased Nrf2 and Nrf2-regulated enzymes’ mRNA expression in brain tissue. The concentrations of superoxide anion radicals, malondialdehyde, and advanced oxidation protein products in brain tissue were consistent with previous results. Overall, our results suggest that NMDARs play a more important role in the pathogenesis of EAE in aged than in young rats. Full article
(This article belongs to the Section Cell Biology and Pathology)
Show Figures

Figure 1

Figure 1
<p>Memantine treatment reduced the mean neurological score and histological score in rats of both age groups. (Panel (<b>A</b>)) Line graph indicates the mean daily neurological score of EAE in rats treated with memantine (+M) or administered with vehicle (−M) for 7 consecutive days from day 7 to day 13 post-immunization. Data are presented as mean ± SEM (data from 4 experiments, n = 6 rats/group/experiment), with statistical differences determined using a two-way ANOVA; # <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; # young vs. aged; * <span class="html-italic">p</span> &lt; 0.05; * +M vs. age-matched −M. (Panel (<b>B</b>)) Representative photomicrographs (<b>a</b>) indicate mononuclear cell infiltrations in spinal cord sections of memantine-treated (+M) or administered with vehicle (−M) young and aged rats. The arrows indicate mononuclear cell infiltrates. Bar graph (<b>b</b>) indicates histological scores for spinal cord of memantine-treated and untreated young and aged EAE rats. Histological scoring is described in the Material and Methods section. Data are presented as mean ± SEM (n = 3 rats/group). * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Memantine reduced the frequency and number of CD4<sup>+</sup> T lymphocytes infiltrating the spinal cord of immunized rats of both age groups. (<b>a</b>) Representative flow cytometry dot plots present CD4 vs. TCRαβ staining of mononuclear cells (MNCs) retrieved at the peak of EAE from spinal cords (SCs) of young and aged rats treated with memantine (+M) or vehicle (−M). Numbers in dot plots represent frequency. Scatter plots present (<b>b</b>) the number of CD4<sup>+</sup> TCRαβ<sup>+</sup> cells and (<b>c</b>) their frequency among MNC cells retrieved from SCs. Data are presented as mean ± SEM from 6 rats per group, with statistical differences determined using a two-way ANOVA. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Memantine reduced reactivation and increased apoptosis of CD4<sup>+</sup> T lymphocytes in the SC more effectively in aged than in young EAE rats. Representative flow cytometry dot plots show (<b>a</b>) CD134 (Panel (<b>A</b>)) and (<b>a</b>) Annexin V (Panel (<b>B</b>)) staining of cells gated as CD4<sup>+</sup> T lymphocytes among mononuclear cells (MNCs) retrieved at the peak of EAE from spinal cords (SCs) of young and aged rats treated with memantine (+M) or vehicle (−M). Numbers in dot plots represent frequency. Scatter plots present the frequency of (<b>b</b>) CD134<sup>+</sup> (Panel (<b>A</b>)) and (<b>b</b>) Annexin V<sup>+</sup> (Panel (<b>B</b>)) cells among CD4<sup>+</sup> T cells in SCs from young and aged +M and −M EAE rats. Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for the frequency of CD134<sup>+</sup> cells (F<sub>(1,20)</sub> = 63.22; <span class="html-italic">p</span> &lt; 0.001) and Annexin V<sup>+</sup> cells (F<sub>(1,20)</sub> = 34.34; <span class="html-italic">p</span> &lt; 0.001). ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>NMDAR antagonist induced lower proportions of IL-17<sup>+</sup> and IFN-γ<sup>+</sup> cells among CD4<sup>+</sup> T cells in SCs of aged rats. Representative flow cytometry dot plots show (<b>a</b>) IL-17 (Panel (<b>A</b>)) and (<b>a</b>) IFN-γ (Panel (<b>B</b>)) staining of cells gated as CD4<sup>+</sup> T lymphocytes among mononuclear cells (MNCs) retrieved from spinal cords (SCs) of young and aged rats treated with memantine (+M) or vehicle (−M) at the peak of EAE. Numbers in dot plots represent frequency. Scatter plots present the frequency of (<b>b</b>) IL-17<sup>+</sup> (Panel (<b>A</b>)) and (<b>b</b>) IFN-γ<sup>+</sup> (Panel (<b>B</b>)) cells among CD4<sup>+</sup> T cells in SCs from young and aged +M and −M EAE rats. Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for the frequency of IL-17<sup>+</sup> cells (F<sub>(1,20)</sub> = 12.88; <span class="html-italic">p</span> &lt; 0.01) and IFN-γ<sup>+</sup> cells (F<sub>(1,20)</sub> = 20.45; <span class="html-italic">p</span> &lt; 0.001). * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Memantine increased the proportion of CX3CR1-expressing microglia more effectively in aged than in young EAE rats. Representative flow cytometry density plot panels present (<b>a</b>) the frequency of CX3CR1<sup>+</sup> cells among CD11b<sup>+</sup>CD45<sup>lo/int</sup> microglia retrieved from spinal cord (SC) mononuclear cells (MNCs) at the peak of EAE from young and aged rats treated with memantine (+M) or vehicle (−M). Numbers in dot plots represent frequency. FMO control incubated with secondary antibody (Ab) alone, without anti-CX3CR1 Ab (-CX3CR1 Ab), was used to settle the gating boundary for CX3CR1<sup>+</sup> cells. (<b>b</b>) Scatter plots indicate the frequency of CX3CR1<sup>+</sup> cells among CD11b<sup>+</sup>CD45<sup>lo/int</sup> microglia. Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for CX3CR1<sup>+</sup> microglial cells (F<sub>(1,20)</sub> = 4.58; <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.</p>
Full article ">Figure 6
<p>Memantine reduced the production of nitric oxide by macrophages/microglia obtained from the spinal cord of aged EAE rats. Bar graphs indicate concentration of nitric oxide in overnight mononuclear spinal cord (SC) cell cultures (<b>a</b>) in the medium alone or (<b>b</b>) stimulated with lipopolysaccharide (LPS). Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for NO concentration determined in LPS-stimulated SC mononuclear cell cultures (F<sub>(1,20)</sub> = 30.50; <span class="html-italic">p</span> &lt; 0.001). *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Memantine was more efficient in reducing oxidative brain tissue damage in aged rats. Bar graphs indicate levels of (<b>a</b>) O2˙<sup>−</sup>, (<b>b</b>) malondialdehyde (MDA), and (<b>c</b>) advanced oxidation protein products (AOPPs) at the peak of EAE in the brain tissue of young and aged EAE rats treated with memantine (+M) or vehicle (−M). Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for the level of AOPP (F<sub>(1,20)</sub> = 12.19; <span class="html-italic">p</span> &lt; 0.01). * <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.</p>
Full article ">Figure 8
<p>Memantine increased the expression of mRNAs for Nrf2 and Nrf2-regulated enzymes. Bar graphs show the fold change in (<b>a</b>) nuclear factor (erythroid-derived 2)-like 2 (Nrf2), (<b>b</b>) heme oxygenase 1 (HO-1), (<b>c</b>) superoxide dismutase (SOD)1, and (<b>d</b>) SOD2 mRNA expression in brain tissue at the peak of EAE from young and aged rats treated with memantine (+M) or vehicle (−M), as determined by qRT-PCR. Results are normalized to β-actin expression. Data are presented as mean ± SEM from 6 rats per group. Two-way ANOVA showed significant interaction between the effects of treatment and age for Nrf2 (F<sub>(1,20)</sub> = 127.0; <span class="html-italic">p</span> &lt; 0.001), HO-1 (F<sub>(1,20)</sub> = 80.06; <span class="html-italic">p</span> &lt; 0.001), SOD1 (F<sub>(1,20)</sub> = 37.91; <span class="html-italic">p</span> &lt; 0.001), and SOD2 (F<sub>(1,20)</sub> = 22.45; <span class="html-italic">p</span> &lt; 0.001) mRNA expression. * <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.</p>
Full article ">
11 pages, 1934 KiB  
Brief Report
Locally Synthetized 17-β-Estradiol Reverses Amyloid-β-42-Induced Hippocampal Long-Term Potentiation Deficits
by Laura Bellingacci, Jacopo Canonichesi, Miriam Sciaccaluga, Alfredo Megaro, Petra Mazzocchetti, Michela Di Mauro, Cinzia Costa, Massimiliano Di Filippo, Vito Enrico Pettorossi and Alessandro Tozzi
Int. J. Mol. Sci. 2024, 25(3), 1377; https://doi.org/10.3390/ijms25031377 - 23 Jan 2024
Cited by 1 | Viewed by 1035
Abstract
Amyloid beta 1-42 (Aβ42) aggregates acutely impair hippocampal long-term potentiation (LTP) of synaptic transmission, and 17β-estradiol is crucial for hippocampal LTP. We tested whether boosting the synthesis of neural-derived 17β-estradiol (nE2) saves hippocampal LTP by the neurotoxic action of Aβ42. Electrophysiological recordings were [...] Read more.
Amyloid beta 1-42 (Aβ42) aggregates acutely impair hippocampal long-term potentiation (LTP) of synaptic transmission, and 17β-estradiol is crucial for hippocampal LTP. We tested whether boosting the synthesis of neural-derived 17β-estradiol (nE2) saves hippocampal LTP by the neurotoxic action of Aβ42. Electrophysiological recordings were performed to measure dentate gyrus (DG) LTP in rat hippocampal slices. Using a pharmacological approach, we tested the ability of nE2 to counteract the LTP impairment caused by acute exposure to soluble Aβ42 aggregates. nE2 was found to be required for LTP in DG under physiological conditions. Blockade of steroid 5α-reductase with finasteride, by increasing nE2 synthesis from testosterone (T), completely recovered LTP in slices treated with soluble Aβ42 aggregates. Modulation of the glutamate N-methyl-D aspartate receptor (NMDAR) by memantine effectively rescued the LTP deficit observed in slices exposed to Aβ42, and memantine prevented LTP reduction observed under the blocking of nE2 synthesis. nE2 is able to counteract Aβ42-induced synaptic dysfunction. This effect depends on a rapid, non-genomic mechanism of action of nE2, which may share a common pathway with glutamate NMDAR signaling. Full article
(This article belongs to the Section Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Effect of modulation of nE2 signaling on DG LTP. (<b>a</b>) Time-course plot of the mean PS amplitude before and after the HFS protocol to induce LTP in control slices (Ctrl) and in the presence of 100 nM letrozole (Let) or 1 nM finasteride (Fin), (Ctrl, n = 6 vs. Let, n = 5, F<sub>(1,9)</sub> = 12.95, ** <span class="html-italic">p</span> &lt; 0.01; Ctrl vs. Fin, n = 5, <span class="html-italic">p</span> &gt; 0.05). Representative pre- and post-HFS traces from a control or a Let-treated slice. (<b>b</b>) Time-course plot of the mean PS amplitude before and after the HFS protocol in Aβ42-treated slices after exposure to 1 nM Fin or co-exposure to Fin plus 100 nM ICI 182,780 (ICI) (Aβ42, n = 8 vs. Aβ42 + Fin, n = 5, F<sub>(1,11)</sub> = 10.2, ** <span class="html-italic">p</span> &lt; 0.01; Aβ42 vs. Aβ42 + Fin + ICI, n = 5, <span class="html-italic">p</span> &gt; 0.05). Representative pre- and post-HFS traces from Aβ42-treated slices in the absence (Ctrl) or presence of Fin. Scale bars: 1 mV, 10 ms.</p>
Full article ">Figure 2
<p>Effect of exogenous 17-β-estradiol (E2) on DG LTP in Aβ42-treated slices. Time-course plot of the mean PS amplitude before and after the HFS protocol to induce LTP in control slices (Ctrl) and in the presence of 1 nM E2 (n = 6).</p>
Full article ">Figure 3
<p>Effect of memantine (Mem) on DG LTP in untreated slices. Time-course plot of the mean PS amplitude before and after the HFS protocol to induce LTP in control slices (Ctrl) and in the presence of 1 μM Mem (n = 6).</p>
Full article ">Figure 4
<p>Effect of memantine (Mem) on DG LTP. (<b>a</b>) Time-course plot of the mean PS amplitude before and after the HFS protocol in Aβ42 plus 1μM Mem-treated slices (Aβ42 + Mem, 253.56 ± 42.54%, n = 7, vs. Aβ42, dashed time-course (Ctrl), F<sub>(1,13)</sub> = 4.78, * <span class="html-italic">p</span> &lt; 0.05). Representative pre- and post-HFS traces from an Aβ42-treated slice in the presence of Mem. (<b>b</b>) Time-course plot of the mean PS amplitude before and after the HFS protocol after the co-exposure of the slices to 100 nM Let plus 1 μM Mem (Mem + Let, 222.80 ± 19.78%, n = 6, vs. Ctrl, dashed time-course, <span class="html-italic">p</span> &gt; 0.05). Representative pre- and post-HFS traces from control slices in the presence of Mem plus Let. Scale bars: 1 mV, 10 ms.</p>
Full article ">Figure 5
<p>Neural E2 modulation of DG LTP in the model of cerebral amyloidosis. Neural 17-β-estradiol (nE2) levels are increased by finasteride (Fin), an inhibitor of the steroid 5α-reductase (5α-RED) that is responsible for dihydrotestosterone (DHT) production from testosterone (T). T is converted to E2 by P450 aromatase (P450 ARO). In hippocampal slices treated for 2 h with 200 nM of the β-amyloid 1-42 fragment (Aβ42), the high-frequency stimulation protocol is able to induce long-term potentiation (LTP) when the slices of the pathological model are treated with Fin or in the presence of 1 μM of the NMDAR antagonist memantine (Mem). Designed with CorelDRAW Graphics Suite.</p>
Full article ">
16 pages, 5203 KiB  
Article
Glutamate’s Effects on the N-Methyl-D-Aspartate (NMDA) Receptor Ion Channel in Alzheimer’s Disease Brain: Challenges for PET Radiotracer Development for Imaging the NMDA Ion Channel
by Nehal M. Shah, Nane Ghazaryan, Noresa L. Gonzaga, Cayz G. Paclibar, Agnes P. Biju, Christopher Liang and Jogeshwar Mukherjee
Molecules 2024, 29(1), 20; https://doi.org/10.3390/molecules29010020 - 19 Dec 2023
Cited by 2 | Viewed by 1228
Abstract
In an effort to further understand the challenges facing in vivo imaging probe development for the N-methyl-D-aspartate (NMDA) receptor ion channel, we have evaluated the effect of glutamate on the Alzheimer’s disease (AD) brain. Human post-mortem AD brain slices of the frontal cortex [...] Read more.
In an effort to further understand the challenges facing in vivo imaging probe development for the N-methyl-D-aspartate (NMDA) receptor ion channel, we have evaluated the effect of glutamate on the Alzheimer’s disease (AD) brain. Human post-mortem AD brain slices of the frontal cortex and anterior cingulate were incubated with [3H]MK-801 and adjacent sections were tested for Aβ and Tau. The binding of [3H]MK-801 was measured in the absence and presence of glutamate and glycine. Increased [3H]MK-801 binding in AD brains was observed at baseline and in the presence of glutamate, indicating a significant increase (>100%) in glutamate-induced NMDA ion channel activity in AD brains compared to cognitively normal brains. The glycine effect was lower, suggesting a decrease of the co-agonist effect of glutamate and glycine in the AD brain. Our preliminary findings suggest that the targeting of the NMDA ion channel as well as the glutamate site may be appropriate in the diagnosis and treatment of AD. However, the low baseline levels of [3H]MK-801 binding in the frontal cortex and anterior cingulate in the absence of glutamate and glycine indicate significant hurdles for in vivo imaging probe development and validation. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of MK-801 and memantine binding to the NMDA ion channel: (<b>A</b>,<b>B</b>). MK-801 with .mol structure (arrow showing nitrogen atom in dark blue) docked to human GluN1/GluN2A NMDA receptor (PDB 6IRH) using Chimera AutoDock. Binding energy for MK-801 was −7.3 Kcal/mol. (<b>C</b>–<b>E</b>). Memantine with .mol structure (arrow showing nitrogen atom in dark blue) docked to human GluN1/GluN2A NMDA receptor (PDB 6IRH) using Chimera AutoDock. Preferred binding site of memantine was somewhat higher than MK-801 with binding energy of −6.7 Kcal/mol (inset shows binding of memantine at the same MK-801 site with binding energy of −5.6 Kcal/mol).</p>
Full article ">Figure 2
<p>Structure of NMDA receptor: (<b>A</b>–<b>C</b>) human GluN1/GluN2A NMDA receptor (PDB 6IRH) complex molecular model illustrating normal ion channel activity and binding of MK-801 in the ion channel (green); (<b>D</b>–<b>F</b>) human GluN1/GluN2A NMDA receptor (PDB 6IRH) complex model in the presence of AD pathology (Aβ plaques and Tau). For simplicity, Aβ fibril is shown interacting with the ion channel to cause it to remain open for extended periods, enabling a greater influx of calcium ions. The ion channel thus has more available binding sites for MK-801 and memantine.</p>
Full article ">Figure 3
<p>Post-mortem human AD brain, frontal cortex section: (<b>A</b>). AD brain frontal cortex slice showing gray (GM) and white matter (WM); (<b>B</b>). Anti-Aβ immunostain of adjacent section showing (arrow in GM) Aβ plaques; (<b>C</b>). Anti-Tau immunostain of adjacent section showing (arrow in GM) abundant Tau; (<b>D</b>). [<sup>3</sup>H]PIB binding to Aβ plaques (arrow) in GM regions [<a href="#B31-molecules-29-00020" class="html-bibr">31</a>]; (<b>E</b>). [<sup>125</sup>I]IPPI binding to Tau (arrow) in GM regions [<a href="#B31-molecules-29-00020" class="html-bibr">31</a>]; (<b>F</b>). Baseline (no drug) [<sup>3</sup>H]MK-801 adjacent to brain slice; (<b>G</b>). [<sup>3</sup>H]MK801 binding, in the presence of 10 μM glutamate, to adjacent slice, showing increased binding to GM regions; (<b>H</b>). [<sup>3</sup>H]MK801 binding, in the presence of 10 μM glycine, to adjacent slice, showing binding to GM regions; (<b>I</b>). [<sup>3</sup>H]MK801 binding, in the presence of 10 μM glutamate and 10 μM glycine, to adjacent slice, showing increased binding to GM regions.</p>
Full article ">Figure 4
<p>Post-mortem human AD and CN brain section analysis: (<b>A</b>). Plot of [<sup>3</sup>H]MK801 binding of all AD and control subjects’ GM frontal cortex regions under different drug conditions, Glu: glutamate; Gly: glycine (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01); (<b>B</b>). Plot of [<sup>3</sup>H]PIB in AD and control subject frontal cortex, showing high levels of Aβ plaques in AD [<a href="#B31-molecules-29-00020" class="html-bibr">31</a>]; (<b>C</b>). Plot of [<sup>125</sup>I]IPPI in AD and control subject frontal cortex, showing high levels of Tau in AD [<a href="#B31-molecules-29-00020" class="html-bibr">31</a>].</p>
Full article ">Figure 5
<p>Post-mortem human AD brain, anterior cingulate section: (<b>A</b>). CN brain, anterior cingulate brain slice showing gray (GM) and white matter (WM); (<b>B</b>–<b>E</b>). [<sup>3</sup>H]MK801 binding in baseline (no drug), 10 μM glycine, 10 μM glutamate and in the presence of 10 μM glycine + 10 μM glutamate to adjacent slice, showing increased binding to GM regions; (<b>F</b>). AD brain, anterior cingulate brain slice showing gray (GM) and white matter (WM); (<b>G</b>–<b>J</b>). [<sup>3</sup>H]MK801 binding in baseline (no drug), 10 μM glycine, 10 μM glutamate and in the presence of 10 μM glycine + 10 μM glutamate to adjacent slice, showing increased binding to GM regions; (<b>K</b>). [<sup>18</sup>F]Flotaza binding to Aβ plaques in GM regions [<a href="#B33-molecules-29-00020" class="html-bibr">33</a>]; (<b>L</b>). Anti-Aβ immunostain of adjacent section showing Aβ plaques; (<b>M</b>). [<sup>124</sup>I]IPPI binding to Tau in GM regions [<a href="#B34-molecules-29-00020" class="html-bibr">34</a>]; (<b>N</b>). Anti-Tau immunostain of adjacent section showing abundant Tau; (<b>O</b>). Plot showing increased GM binding of [<sup>3</sup>H]MK-801, Glu + Gly &gt; Glu &gt; Gly &gt; baseline, to available NMDA ion channel sites under different drug conditions in CN and AD brains.</p>
Full article ">Figure 6
<p>Chemical structures of [<sup>3</sup>H]MK-801 (<b>1</b>), PET radiotracers based on MK801 (<b>2</b>–<b>5</b>) and SPECT radiotracer based on MK-801 (<b>6</b>). Chemical structures of PET radiotracers based on PCP and TCP (<b>7</b>–<b>12</b>).</p>
Full article ">Figure 7
<p>Radiolabeled carbon-11 (<b>13</b>,<b>16</b>), fluorine-18 (<b>15</b>) and iodine-123 (<b>14</b>) guanidine analogs investigated for the NMDA ion channel.</p>
Full article ">Figure 8
<p>Chemical structures of GluN2B, showing ifenprodil (<b>17</b>) and the fluorine-18 (<b>18</b>) and carbon-11 (<b>19</b>) PET radiotracers derived from it. The related benzazepine derivative [<sup>18</sup>F]<b>20</b> has been successfully used for PET imaging studies for the GluN2B subunit.</p>
Full article ">Figure 9
<p>Rat brain [<sup>3</sup>H]MK-801 and [<sup>18</sup>F]Mefway comparison: (<b>A</b>). Normal rat brain, horizontal brain slice showing baseline (no drug) binding of [<sup>3</sup>H]MK-801; (<b>B</b>). Normal rat brain, horizontal brain slice showing binding of [<sup>3</sup>H]MK-801 in the presence of glutamate (Glu, 10 μM) and glycine (Gly, 10 μM); (<b>C</b>). Normal rat brain, horizontal brain slice showing baseline (no drug) binding of [<sup>18</sup>F]Mefway to serotonin 5-HT1A receptors; (<b>D</b>). Scan of normal rat brain, horizontal brain slice 10 μm thick; (<b>E</b>). Plot showing increased binding of [<sup>3</sup>H]MK-801 to available NMDA ion channel sites in the presence of Glu + Gly, compared to baseline; (<b>F</b>). Plot showing ratio of brain regions versus cerebellum for [<sup>3</sup>H]MK-801 binding to NMDA ion channels and [<sup>18</sup>F]Mefway to serotonin 5-HT1A receptors (FC: frontal cortex; AC: anterior cingulate; ST: striatum; TH: thalamus; HP: hippocampus; CB: cerebellum).</p>
Full article ">Figure 10
<p>Potential strategy for in vivo radiotracer development for the NMDA receptor ion channel. (<b>A</b>). In vitro phase of radiotracer development including demonstration of binding sensitivity to glutamate and glycine-induced increased availability of ion-channels; (<b>B</b>). In vivo preclinical phase of testing the radiotracer in animal model studies with known brain fluctuations in glutamate and glycine concentrations; (<b>C</b>). Translation of selected radiotracer to human feasibility studies to assess increased “availability” of NMDA receptor ion-channels.</p>
Full article ">
16 pages, 2825 KiB  
Article
Inhibition of NMDA Receptor Activation in the Rostral Ventrolateral Medulla by Amyloid-β Peptide in Rats
by Md Sharyful Islam, Chih-Chia Lai, Lan-Hui Wang and Hsun-Hsun Lin
Biomolecules 2023, 13(12), 1736; https://doi.org/10.3390/biom13121736 - 2 Dec 2023
Cited by 2 | Viewed by 1791
Abstract
N-methyl-D-aspartate (NMDA) receptors, a subtype of ionotropic glutamate receptors, are important in regulating sympathetic tone and cardiovascular function in the rostral ventrolateral medulla (RVLM). Amyloid-beta peptide (Aβ) is linked to the pathogenesis of Alzheimer’s disease (AD). Cerebro- and cardiovascular diseases might be the [...] Read more.
N-methyl-D-aspartate (NMDA) receptors, a subtype of ionotropic glutamate receptors, are important in regulating sympathetic tone and cardiovascular function in the rostral ventrolateral medulla (RVLM). Amyloid-beta peptide (Aβ) is linked to the pathogenesis of Alzheimer’s disease (AD). Cerebro- and cardiovascular diseases might be the risk factors for developing AD. The present study examines the acute effects of soluble Aβ on the function of NMDA receptors in rats RVLM. We used the magnitude of increases in the blood pressure (pressor responses) induced by microinjection of NMDA into the RVLM as an index of NMDA receptor function in the RVLM. Soluble Aβ was applied by intracerebroventricular (ICV) injection. Aβ1-40 at a lower dose (0.2 nmol) caused a slight reduction, and a higher dose (2 nmol) showed a significant decrease in NMDA-induced pressor responses 10 min after administration. ICV injection of Aβ1-42 (2 nmol) did not affect NMDA-induced pressor responses in the RVLM. Co-administration of Aβ1-40 with ifenprodil or memantine blocked the inhibitory effects of Aβ1-40. Immunohistochemistry analysis showed a significant increase in the immunoreactivity of phosphoserine 1480 of GluN2B subunits (pGluN2B-serine1480) in the neuron of the RVLM without significant changes in phosphoserine 896 of GluN1 subunits (pGluN1-serine896), GluN1 and GluN2B, 10 min following Aβ1-40 administration compared with saline. Interestingly, we found a much higher level of Aβ1-40 compared to that of Aβ1-42 in the cerebrospinal fluid (CSF) measured using enzyme-linked immunosorbent assay 10 min following ICV administration of the same dose (2 nmol) of the peptides. In conclusion, the results suggest that ICV Aβ1-40, but not Aβ1-42, produced an inhibitory effect on NMDA receptor function in the RVLM, which might result from changes in pGluN2B-serine1480 (regulated by casein kinase II). The different elimination of the peptides in the CSF might contribute to the differential effects of Aβ1-40 and Aβ1-42 on NMDA receptor function. Full article
(This article belongs to the Special Issue NMDA Receptor in Health and Diseases: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Effect of intracerebroventricular (i.c.v.) injection of amyloid-beta (Aβ) on NMDA receptor function in the RVLM. Representative recordings showed the effects of i.c.v. administration Aβ1-40 (0.2 nmol) (<b>A</b>), Aβ1-40 (2 nmol) (<b>B</b>), and Aβ1-42 (2 nmol) (<b>C</b>) on pulsatile arterial pressure (PAP) and mean arterial pressure (MAP) induced by microinjection of NMDA (0.14 nmol) into the RVLM every 30 min. (<b>D</b>) The bar graph shows the percentage changes in NMDA-induced increases in MAP 20 min before (control, i.e., 100%) and 10, 40 min after i.c.v. injection of Aβ1-40 (0.2 nmol, <span class="html-italic">n</span> = 12), (2 nmol, <span class="html-italic">n</span> = 14), and Aβ1-42 (2 nmol, <span class="html-italic">n</span> = 9). Values are mean + SEM. ** <span class="html-italic">p</span> ≤ 0.01 compared to the corresponding control analyzed using one-way ANOVA repeated measures followed by Bonferroni’s multiple comparison post-test.</p>
Full article ">Figure 2
<p>Effect of intracerebroventricular (i.c.v.) co-injection of amyloid-beta (Aβ) with ifenprodil, a GluN2B receptor antagonist, on NMDA receptor function in the RVLM. Representative recordings showed the effects of i.c.v. administration of ifenprodil alone at doses of 0.5 nmol (<b>A</b>) and 2.5 nmol (<b>B</b>), co-administration of ifenprodil with Aβ1-40 (<b>C</b>) and co-administration of ifenprodil with Aβ1-42 (<b>D</b>) on increases in pulsatile arterial pressure (PAP) and mean arterial pressure (MAP) induced by microinjection of NMDA (0.14 nmol) into the RVLM. NMDA was applied every 30 min. (<b>E</b>) The bar graph shows the percentage changes in NMDA-induced increases in MAP by i.c.v. injection of ifenprodil (0.5 nmol, <span class="html-italic">n</span> = 5 and 2.5 nmol, <span class="html-italic">n</span> = 5), (<b>F</b>) co-administration of Aβ1-40 (2 nmol) with ifenprodil (0.5 nmol), <span class="html-italic">n</span> = 6, and (<b>G</b>) co-administration of Aβ1-42 (2 nmol) with ifenprodil (0.5 nmol), <span class="html-italic">n</span> = 6, 20 min before (control, i.e., 100%) and 10, 40 min after application of NMDA. Values are mean + SEM. * <span class="html-italic">p</span> ≤ 0.05, and ** <span class="html-italic">p</span> ≤ 0.01 compared to the corresponding control analyzed using one-way ANOVA repeated measures followed by Bonferroni’s multiple comparison post-test.</p>
Full article ">Figure 3
<p>Effect of intracerebroventricular (i.c.v.) co-injection of amyloid-beta (Aβ) with memantine, an NMDA channel blocker, on NMDA receptor function in the RVLM. Representative recordings showed the effects of i.c.v. administration of memantine (9 nmol, <span class="html-italic">n</span> = 6) alone (<b>A</b>), co-administration of memantine with Aβ1-40 (<b>B</b>), and co-administration of memantine with Aβ1-42 (<b>C</b>) on increases in pulsatile arterial pressure (PAP) and mean arterial pressure (MAP) induced by microinjection of NMDA (0.14 nmol) into the RVLM. NMDA was applied every 30 min. (<b>D</b>) The bar graph shows the percentage changes in NMDA-induced increases in MAP by i.c.v. injection of memantine (9 nmol, <span class="html-italic">n</span> = 6), (<b>E</b>) co-administration of Aβ1-40 (2 nmol) with memantine (9 nmol), <span class="html-italic">n</span> = 4 and (<b>F</b>) co-administration of Aβ1-42 (2 nmol) with memantine (9 nmol) <span class="html-italic">n</span> = 6, 20 min before (control, i.e., 100%) and 10, 40 min after application of NMDA. Values are mean + SEM. ** <span class="html-italic">p</span> ≤ 0.01 compared to the corresponding control analyzed using one-way ANOVA repeated measures followed by Bonferroni’s multiple comparison post-test.</p>
Full article ">Figure 4
<p>Representative immunohistochemical images show immunoreactive GluN1 (<b>A</b>, <b>left</b>) or pGluN1-serine 896 (<b>B</b>, <b>left</b>) neurons in the ventral brainstem at low or high magnification 10 min after saline (5 μL) or Aβ1-40 (2 nmol/5 μL) ICV treatment. The red box refers to the RVLM. The scale bar is 500 μm. The bar graph showed the number of immunoreactive GluN1 (<b>A</b>, <b>right</b>) or pGluN1-serine 896 (<b>B</b>, <b>right</b>) neurons in the RVLM after saline or Aβ1-40 treatment. Values are mean + SEM.</p>
Full article ">Figure 5
<p>Representative immunofluorescence (IF) images show immunoreactive GluN2B (<b>A</b>, <b>left</b>) or pGluN2B-serine1480 (<b>B</b>, <b>left</b>) neurons in the ventral brainstem 10 min after saline (5 μL) or Aβ1-40 (2 nmol/5 μL) ICV treatment. The red box refers to the RVLM. The scale bar is 200 μm. The bar graphs showed the number of GluN2B-IF (<b>A</b>, <b>right</b>) or pGluN2B-serine1480-IF (<b>B</b>, <b>right</b>) neurons in the RVLM after saline or Aβ1-40 treatment. Values are mean + SEM. ** <span class="html-italic">p</span> ≤ 0.01 compared to the saline group analyzed using unpaired t-test.</p>
Full article ">Figure 6
<p>The graph shows the concentration of Aβ1-42 and Aβ1-40 in cerebrospinal fluid 10 min after intracerebroventricular injection of both peptides. Values are mean ± SEM. <span class="html-italic">n</span> = 5 for each group. There is a significant difference between the two groups analyzed using unpaired <span class="html-italic">t</span>-test.</p>
Full article ">
13 pages, 1999 KiB  
Article
In Vitro Permeability Study of Homotaurine Using a High-Performance Liquid Chromatography with Fluorescence Detection Pre-Column Derivatization Method
by Marianna Ntorkou, Eleni Tsanaktsidou, Konstantina Chachlioutaki, Dimitrios G. Fatouros and Catherine K. Markopoulou
Molecules 2023, 28(20), 7086; https://doi.org/10.3390/molecules28207086 - 14 Oct 2023
Cited by 1 | Viewed by 1272
Abstract
Homotaurine (HOM) is considered a promising drug for the treatment of Alzheimer’s and other neurodegenerative diseases. In the present work, a new high-performance liquid chromatography with fluorescence detection (HPLC–FLD) (λex. = 340 nm and λem. = 455 nm) method was developed and validated [...] Read more.
Homotaurine (HOM) is considered a promising drug for the treatment of Alzheimer’s and other neurodegenerative diseases. In the present work, a new high-performance liquid chromatography with fluorescence detection (HPLC–FLD) (λex. = 340 nm and λem. = 455 nm) method was developed and validated for the study of substance permeability in the central nervous system (CNS). Analysis was performed on a RP-C18 column with a binary gradient elution system consisting of methanol–potassium phosphate buffer solution (pH = 7.0, 0.02 M) as mobile phase. Samples of homotaurine and histidine (internal standard) were initially derivatized with ortho-phthalaldehyde (OPA) (0.01 M), N-acetylcysteine (0.01 M) and borate buffer (pH = 10.5; 0.05 M). To ensure the stability and efficiency of the reaction, the presence of different nucleophilic reagents, namely (a) 2-mercaptoethanol (2-ME), (b) N-acetylcysteine (NAC), (c) tiopronin (Thiola), (d) 3-mercaptopropionic acid (3-MPA) and (e) captopril, was investigated. The method was validated (R2 = 0.9999, intra-day repeatability %RSD < 3.22%, inter-day precision %RSD = 1.83%, limits of detection 5.75 ng/mL and limits of quantification 17.43 ng/mL, recovery of five different concentrations 99.75–101.58%) and successfully applied to investigate the in vitro permeability of homotaurine using Franz diffusion cells. The apparent permeability (Papp) of HOM was compared with that of memantine, which is considered a potential therapeutic drug for various CNSs. Our study demonstrates that homotaurine exhibits superior permeability through the simulated blood–brain barrier compared to memantine, offering promising insights for enhanced drug delivery strategies targeting neurological conditions. Full article
(This article belongs to the Section Analytical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Stability study of homotaurine derivative in the presence of: (<b>a</b>) 2-ME; (<b>b</b>) 3-MPA; (<b>c</b>) NAC.</p>
Full article ">Figure 1 Cont.
<p>Stability study of homotaurine derivative in the presence of: (<b>a</b>) 2-ME; (<b>b</b>) 3-MPA; (<b>c</b>) NAC.</p>
Full article ">Figure 2
<p>Stability study of homotaurine and histidine derivatives in the presence of NAC.</p>
Full article ">Figure 3
<p>Derivatization reaction of homotaurine in presence of NAC.</p>
Full article ">Figure 4
<p>Representative chromatogram of: (<b>a</b>) blank of the reaction; (<b>b</b>) blank of PBS matrix and (<b>c</b>) spiked sample with homotaurine (45.5 ng/mL, t<sub>R</sub> = 6.034 s) and histidine (45.5 ng/mL, t<sub>R</sub> = 8.850 s). Red arrows indicate the time points at which the histidine and homotaurine derivatives are eluted, at 6.034 and 8.850 s, respectively. At these time points, there are no interfering peaks at the two blank chromatograms.</p>
Full article ">Figure 5
<p>(<b>a</b>) Permeability study of homotaurine; (<b>b</b>) Permeability study of memantine.</p>
Full article ">Scheme 1
<p>Chemical structures of taurine and homotaurine.</p>
Full article ">
25 pages, 886 KiB  
Review
Treatment of Alzheimer’s Disease: Beyond Symptomatic Therapies
by Francesca R. Buccellato, Marianna D’Anca, Gianluca Martino Tartaglia, Massimo Del Fabbro, Elio Scarpini and Daniela Galimberti
Int. J. Mol. Sci. 2023, 24(18), 13900; https://doi.org/10.3390/ijms241813900 - 9 Sep 2023
Cited by 20 | Viewed by 5778
Abstract
In an ever-increasing aged world, Alzheimer’s disease (AD) represents the first cause of dementia and one of the first chronic diseases in elderly people. With 55 million people affected, the WHO considers AD to be a disease with public priority. Unfortunately, there are [...] Read more.
In an ever-increasing aged world, Alzheimer’s disease (AD) represents the first cause of dementia and one of the first chronic diseases in elderly people. With 55 million people affected, the WHO considers AD to be a disease with public priority. Unfortunately, there are no final cures for this pathology. Treatment strategies are aimed to mitigate symptoms, i.e., acetylcholinesterase inhibitors (AChEI) and the N-Methyl-D-aspartate (NMDA) antagonist Memantine. At present, the best approaches for managing the disease seem to combine pharmacological and non-pharmacological therapies to stimulate cognitive reserve. Over the last twenty years, a number of drugs have been discovered acting on the well-established biological hallmarks of AD, deposition of β-amyloid aggregates and accumulation of hyperphosphorylated tau protein in cells. Although previous efforts disappointed expectations, a new era in treating AD has been working its way recently. The Food and Drug Administration (FDA) gave conditional approval of the first disease-modifying therapy (DMT) for the treatment of AD, aducanumab, a monoclonal antibody (mAb) designed against Aβ plaques and oligomers in 2021, and in January 2023, the FDA granted accelerated approval for a second monoclonal antibody, Lecanemab. This review describes ongoing clinical trials with DMTs and non-pharmacological therapies. We will also present a future scenario based on new biomarkers that can detect AD in preclinical or prodromal stages, identify people at risk of developing AD, and allow an early and curative treatment. Full article
(This article belongs to the Special Issue Basic, Translational and Clinical Research on Dementia)
Show Figures

Figure 1

Figure 1
<p>Summary of Alzheimer’s disease (AD) treatment strategies in clinical trials. Created with <a href="http://Biorender.com" target="_blank">Biorender.com</a>.</p>
Full article ">
14 pages, 3303 KiB  
Article
Development of an In Vitro Methodology to Assess the Bioequivalence of Orally Disintegrating Tablets Taken without Water
by Toshihide Takagi, Takato Masada, Keiko Minami, Makoto Kataoka and Shinji Yamashita
Pharmaceutics 2023, 15(9), 2192; https://doi.org/10.3390/pharmaceutics15092192 - 24 Aug 2023
Cited by 3 | Viewed by 1589
Abstract
To assess the probability of bioequivalence (BE) between orally disintegrating tablets (ODTs) taken without water and conventional tablets (CTs) taken with water, an in vitro biorelevant methodology was developed using the BE Checker, which reproduces fluid shifts in the gastrointestinal tract and drug [...] Read more.
To assess the probability of bioequivalence (BE) between orally disintegrating tablets (ODTs) taken without water and conventional tablets (CTs) taken with water, an in vitro biorelevant methodology was developed using the BE Checker, which reproduces fluid shifts in the gastrointestinal tract and drug permeation. In addition to the fluid shift from the stomach to the small intestine, the process of ODT disintegration in a small amount of fluid in the oral cavity and the difference in gastric emptying caused by differences in water intake were incorporated into the evaluation protocol. Assuming a longer time to maximum plasma concentration after oral administration of ODTs taken without water than for CTs taken with water due to a delay in gastric emptying, the fluid shift in the donor chamber of the BE Checker without water was set longer than that taken with water. In the case of naftopidil ODTs and CTs, the values of the f2 function, representing the similarity of the permeation profiles, were 50 or higher when the fluid shift in ODTs taken without water was set at 1.5 or 2 times longer than that of the CTs taken with water. The values of the f2 function in permeation profiles of pitavastatin and memantine ODTs were both 62 when the optimized experimental settings for naftopidil formulations were applied. This methodology can be useful in formulation studies for estimating the BE probability between ODTs and CTs. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration and pictures of the BE Checker to assess the BE of (<b>A</b>) CTs taken with water and (<b>B</b>) ODTs taken without water.</p>
Full article ">Figure 2
<p>Experimental conditions in the BE Checker for tablets taken with and without water. For (<b>A</b>) CTs taken with water, (<b>B</b>) ODTs taken without water.</p>
Full article ">Figure 3
<p>Simulated pH–time profiles of the donor fluid in the BE Checker.</p>
Full article ">Figure 4
<p>Dissolution and permeation profiles of naftopidil from 25 mg CTs and ODTs in the BE Checker when the stirring rate was 50 rpm. (<b>A</b>) Dissolved amount in the donor chamber; (<b>B</b>) dissolved concentration in the donor chamber; (<b>C</b>) permeated amount in the receiver chamber.</p>
Full article ">Figure 5
<p>Dissolution and permeation profiles of naftopidil from 25 mg CTs and ODTs in the BE Checker when the stirring rate was 100 rpm. (<b>A</b>) Dissolved amount in the donor chamber; (<b>B</b>) dissolved concentration in the donor chamber; (<b>C</b>) permeated amount in the receiver chamber.</p>
Full article ">Figure 6
<p>Dissolution and permeation profiles of pitavastatin from 1 mg CTs and ODTs in the BE Checker when the stirring rate was 100 rpm. (<b>A</b>) Dissolved amount in the donor chamber; (<b>B</b>) dissolved concentration in the donor chamber; (<b>C</b>) permeated amount in the receiver chamber.</p>
Full article ">Figure 7
<p>Dissolution and permeation profiles of memantine from 5 mg CTs and ODTs in the BE Checker when the stirring rate was 100 rpm. (<b>A</b>) Dissolved amount in the donor chamber; (<b>B</b>) dissolved concentration in the donor chamber; (<b>C</b>) permeated amount in the receiver chamber.</p>
Full article ">
16 pages, 1971 KiB  
Article
Memantine Improves the Disturbed Glutamine and γ-Amino Butyric Acid Homeostasis in the Brain of Rats Subjected to Experimental Autoimmune Encephalomyelitis
by Beata Dąbrowska-Bouta, Lidia Strużyńska, Marta Sidoryk-Węgrzynowicz and Grzegorz Sulkowski
Int. J. Mol. Sci. 2023, 24(17), 13149; https://doi.org/10.3390/ijms241713149 - 24 Aug 2023
Cited by 2 | Viewed by 1139
Abstract
Glutamine (Gln), glutamate (Glu), and γ-amino butyric acid (GABA) are essential amino acids for brain metabolism and function. Astrocyte-derived Gln is the precursor for the two most important neurotransmitters in the central nervous system (CNS), which are the excitatory neurotransmitter Glu and the [...] Read more.
Glutamine (Gln), glutamate (Glu), and γ-amino butyric acid (GABA) are essential amino acids for brain metabolism and function. Astrocyte-derived Gln is the precursor for the two most important neurotransmitters in the central nervous system (CNS), which are the excitatory neurotransmitter Glu and the inhibitory neurotransmitter GABA. In addition to their roles in neurotransmission, these amino acids can be used as alternative substrates in brain metabolism that enable metabolic coupling between astrocytes and neurons in the glutamate–glutamine cycle (GGC). The disturbed homeostasis of these amino acids within the tripartite synapse may be involved in the pathogenesis of various neurological diseases. Interactions between astrocytes and neurons in terms of Gln, Glu, and GABA homeostasis were studied in different phases of experimental allergic encephalomyelitis (EAE) in Lewis rats. The results of the study showed a decrease in the transport (uptake and release) of Gln and GABA in both neuronal and astrocyte-derived fractions. These effects were fully or partially reversed when the EAE rats were treated with memantine, a NMDA receptor antagonist. Changes in the expression and activity of selected glutamine/glutamate metabolizing enzymes, such as glutamine synthase (GS) and phosphate-activated glutaminase (PAG), which were affected by memantine, were observed in different phases of EAE. The results suggested perturbed homeostasis of Gln, Glu, and GABA during EAE, which may indicate alterations in neuron–astrocyte coupling and dysfunction of the tripartite synapse. Memantine appears to partially regulate the disturbed relationships between Gln, Glu, and GABA. Full article
(This article belongs to the Special Issue Neurotransmitters in Neurodegenerative Diseases)
Show Figures

Figure 1

Figure 1
<p>Na<sup>+</sup>-dependent [<sup>14</sup> C] GABA uptake into synaptosomes (<b>A</b>) and GPV fraction (<b>B</b>) obtained from control, EAE, and EAE + memantine rat brains during the acute phase of the disease (12 d.p.i.). GABA release from synaptosomes (<b>C</b>) and GPV fraction (<b>D</b>) obtained from control, EAE, and EAE + memantine rat brain tissue during the acute phase of the disease (12 d.p.i.). Bars represent [<sup>14</sup>C]GABA radioactivity released from pellets after depolarization of membranes with 50 mM KCl at a maximum of the uptake curves (4 min). The radioactivity was assayed 6 min after depolarization. The results represent the means ± SD from six separate experiments performed in triplicate; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 significantly different vs. control group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. EAE rats not subjected to therapy with memantine (one-way ANOVA with post hoc Tukey’s test).</p>
Full article ">Figure 2
<p>Na<sup>+</sup>-dependent [<sup>3</sup>H]glutamine uptake into synaptosomes (<b>A</b>) and GPV fraction (<b>B</b>) obtained from the control, EAE, and EAE + memantine rat brain tissue during the acute phase of the disease (12 d.p.i.). Glutamine release from synaptosomes (<b>C</b>) and the GPV fraction (<b>D</b>) obtained from control, EAE, and EAE + memantine rat brain tissue during the acute phase of the disease (12 d.p.i.). Bars represent [<sup>3</sup>H]glutamine radioactivity released from pellets after the depolarization of membranes with 50 mM KCl at a maximum of the uptake curves (5 min). The radioactivity was assayed 6 min after depolarization. Results represent the means ± SD from six separate experiments performed in triplicate; * <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 significantly different vs. control group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 vs. EAE rats not subjected to therapy with memantine (one-way ANOVA with post hoc Tukey’s test).</p>
Full article ">Figure 3
<p>Protein expression of glutamine synthase (GS) (<b>A</b>) and phosphate-activated glutaminase (PAG) (<b>B</b>) in the brain tissue of control and EAE rats measured at different times post-immunization and after therapeutic treatment with memantine. Representative immunoblots and graphs (<b>A</b>,<b>C</b>) showing the results of densitometric analysis of four independent immunoblots, normalized to <span class="html-italic">β</span>-actin, each performed from a distinct rat brain. Expression of GS and PAG mRNAs (<b>B</b>,<b>D</b>) in the forebrains of the control, EAE, and EAE rats after treatment with memantine in different phases of EAE. The mRNA levels were determined by quantitative real-time PCR (see <a href="#sec4-ijms-24-13149" class="html-sec">Section 4</a>) and normalized to actin. The results are the means from <span class="html-italic">n</span> = 6 animals in each group and are expressed as a percentage of the control. * <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 significantly different vs. control rats; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 significantly different vs. EAE rats not subjected to therapy (one-way ANOVA followed by Sidak’s multiple comparison post-test).</p>
Full article ">Figure 4
<p>The activity of the astrocyte-specific enzyme glutamine synthase (GS) (<b>A</b>) and neuron-specific enzyme phosphate-activated glutaminase (PAG) (<b>B</b>) in the brain tissue of control and EAE rats and after therapeutic treatment with memantine at the peak of EAE at 12 d.p.i. All rats in the EAE group without memantine treatment had a neurological symptoms score of 4. The results are the means from <span class="html-italic">n</span> = 6 animals in each group and are expressed as means ± SD. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001 significantly different vs. control rats; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001 significantly different vs. EAE rats not subjected to therapy (one-way ANOVA followed by Tukey’s multiple comparison post-test).</p>
Full article ">
14 pages, 3387 KiB  
Article
Investigation of Cognitive Impairment in the Course of Post-COVID Syndrome
by Milena Dimitrova, Yoanna Marinova and Dancho Dilkov
Diagnostics 2023, 13(16), 2703; https://doi.org/10.3390/diagnostics13162703 - 18 Aug 2023
Viewed by 1311
Abstract
(1) Background: The study presents results from an investigation of cognitive impairment in patients hospitalized in the first psychiatric clinic in Bulgaria to treat patients with COVID-19 during the pandemic period between 2020 and 2022. One hundred and twenty patients who had recovered [...] Read more.
(1) Background: The study presents results from an investigation of cognitive impairment in patients hospitalized in the first psychiatric clinic in Bulgaria to treat patients with COVID-19 during the pandemic period between 2020 and 2022. One hundred and twenty patients who had recovered from acute COVID-19 infection (up to 12 weeks ago) and had no previous history of cognitive impairment participated in the study. In 23 of them (19.17%), disturbance of cognitive functioning was observed. (2) Methods: All 23 patients underwent neuropsychological (Luria’s test, Platonov’s Maze test, MMSE, Boston Naming test) and neuroimaging examinations. Only seven of them had evidence of cortical atrophy on CT/MRI images. The most significantly demonstrative image of one of those patients is presented. (3) Results: The neuropsychological testing results of both groups show a certain decrease in fixation and memory retention as well as in the range, concentration, distribution and switching of attention. Deviations from the norm on the MMSE, as well as on the Boston Naming Test, were found in the group of patients with cortical atrophy (mild to moderate aphasia). Neuroprotective agents such as Citicoline, Piracetam and Memantine were prescribed to the patients with evident cortical atrophy. After 3 months, positive results of the neuropsychological examination were reported in both groups. (4) Conclusions: Although there are limited data on the benefit of prescribing pro-cognitive agents in the post-COVID period, our clinical experience suggests that it might be useful in the recovery process from the infection’s consequences on cognition for patients with brain pathology. Full article
(This article belongs to the Special Issue New Advances in the Diagnosis and Treatment of Mental Disorders)
Show Figures

Figure 1

Figure 1
<p>Neuropsychological testing results of patients with cortical atrophy.</p>
Full article ">Figure 2
<p>General improvement in cognitive functioning of the patients with cortical atrophy 3 months after treatment with pro-cognitive agent.</p>
Full article ">Figure 3
<p>Neuropsychological testing results of patients without cortical atrophy.</p>
Full article ">Figure 4
<p>General improvement in cognitive functioning of the patients without cortical atrophy 3 months after treatment without pro-cognitive agent.</p>
Full article ">Figure 5
<p>Comorbid psychiatric disorders in the 23 patients with objective cognitive decline in the course of post-COVID syndrome.</p>
Full article ">Figure 6
<p>MRI image of the patient conducted in 2020 before the COVID-19 infection.</p>
Full article ">Figure 7
<p>CT scan image of the patient conducted in 2022 after surviving 2 severe COVID-19 infections.</p>
Full article ">Scheme 1
<p>Flow diagram illustrating the methodology of the study (inclusion and exclusion criteria, neuropsychology testing and neuroimaging examinations). CT—computed tomography; MRI—magnetic resonance imaging.</p>
Full article ">
Back to TopTop