[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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (840)

Search Parameters:
Keywords = α-synuclein

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 1078 KiB  
Review
Potential Mechanisms of Tunneling Nanotube Formation and Their Role in Pathology Spread in Alzheimer’s Disease and Other Proteinopathies
by Szymon Kotarba, Marta Kozłowska, Małgorzata Scios, Kamil Saramowicz, Julia Barczuk, Zuzanna Granek, Natalia Siwecka, Wojciech Wiese, Michał Golberg, Grzegorz Galita, Grzegorz Sychowski, Ireneusz Majsterek and Wioletta Rozpędek-Kamińska
Int. J. Mol. Sci. 2024, 25(19), 10797; https://doi.org/10.3390/ijms251910797 - 8 Oct 2024
Viewed by 546
Abstract
Alzheimer’s disease (AD) is the most common type of dementia worldwide. The etiopathogenesis of this disease remains unknown. Currently, several hypotheses attempt to explain its cause, with the most well-studied being the cholinergic, beta-amyloid (Aβ), and Tau hypotheses. Lately, there has been increasing [...] Read more.
Alzheimer’s disease (AD) is the most common type of dementia worldwide. The etiopathogenesis of this disease remains unknown. Currently, several hypotheses attempt to explain its cause, with the most well-studied being the cholinergic, beta-amyloid (Aβ), and Tau hypotheses. Lately, there has been increasing interest in the role of immunological factors and other proteins such as alpha-synuclein (α-syn) and transactive response DNA-binding protein of 43 kDa (TDP-43). Recent studies emphasize the role of tunneling nanotubes (TNTs) in the spread of pathological proteins within the brains of AD patients. TNTs are small membrane protrusions composed of F-actin that connect non-adjacent cells. Conditions such as pathogen infections, oxidative stress, inflammation, and misfolded protein accumulation lead to the formation of TNTs. These structures have been shown to transport pathological proteins such as Aβ, Tau, α-syn, and TDP-43 between central nervous system (CNS) cells, as confirmed by in vitro studies. Besides their role in spreading pathology, TNTs may also have protective functions. Neurons burdened with α-syn can transfer protein aggregates to glial cells and receive healthy mitochondria, thereby reducing cellular stress associated with α-syn accumulation. Current AD treatments focus on alleviating symptoms, and clinical trials with Aβ-lowering drugs have proven ineffective. Therefore, intensifying research on TNTs could bring scientists closer to a better understanding of AD and the development of effective therapies. Full article
Show Figures

Figure 1

Figure 1
<p>Tunneling nanotubes (TNTs) are tubular, membranous structures that contain F-actin. These structures often consist of bundles of individuals (iTNTs), where each iTNT is encased by a plasma membrane and interconnected with others through bridging threads containing N-cadherin. TNTs are classified into two categories based on their diameter: “thin” TNTs and “thick” TNTs. “Thin” TNTs, ranging from 20 to 700 nanometers in diameter, primarily facilitate the exchange of smaller cargo, such as secondary messengers, small peptides, and molecules with a molecular weight below 1.2 kDa. Conversely, “thick” TNTs, which exceed 700 nanometers in diameter, are capable of transporting larger cargo, including organelles, viruses, and molecules larger than 1.2 kDa.</p>
Full article ">Figure 2
<p>Evidence indicates that amyloid beta (Aβ) can move bidirectionally through tunneling nanotubes (TNTs) in various cell lines. In Alzheimer’s disease (AD), excess Aβ released from cells is rapidly transferred to neighboring cells via TNTs, which accelerates disease progression. Tau may also trigger TNT formation, and fibrillar Tau transport through TNTs has been confirmed in vitro. Both exogenous and endogenous Tau aggregates can be transmitted between cells and have been detected inside TNTs in neuronal cell lines. Alpha-synuclein (α-syn) may be transferred via TNTs in vitro, reducing its burden in donor cells. Similar to Tau, α-syn increases the number of TNT connections compared to untreated cells and can be transported between neurons and microglia. α-Syn aggregates are preferentially transferred from neuronal to microglial cells, while mitochondria are transported in the opposite direction. Additionally, in lymphoblasts from AD patients, increased formation of actin protrusions resembling TNTs or TNT-like structures was observed. Transactive response DNA-binding protein of 43 kDa (TDP-43) aggregates were found alongside F-actin fibers in the cytosolic compartment of these cells and were also detected within tubular actin channels.</p>
Full article ">
25 pages, 15275 KiB  
Article
Oral Trehalose Intake Modulates the Microbiota–Gut–Brain Axis and Is Neuroprotective in a Synucleinopathy Mouse Model
by Solène Pradeloux, Katherine Coulombe, Alexandre Jules Kennang Ouamba, Amandine Isenbrandt, Frédéric Calon, Denis Roy and Denis Soulet
Nutrients 2024, 16(19), 3309; https://doi.org/10.3390/nu16193309 - 30 Sep 2024
Viewed by 735
Abstract
Parkinson’s disease (PD) is a neurodegenerative disease affecting dopaminergic neurons in the nigrostriatal and gastrointestinal tracts, causing both motor and non-motor symptoms. This study examined the neuroprotective effects of trehalose. This sugar is confined in the gut due to the absence of transporters, [...] Read more.
Parkinson’s disease (PD) is a neurodegenerative disease affecting dopaminergic neurons in the nigrostriatal and gastrointestinal tracts, causing both motor and non-motor symptoms. This study examined the neuroprotective effects of trehalose. This sugar is confined in the gut due to the absence of transporters, so we hypothesized that trehalose might exert neuroprotective effects on PD through its action on the gut microbiota. We used a transgenic mouse model of PD (PrP-A53T G2-3) overexpressing human α-synuclein and developing GI dysfunctions. Mice were given water with trehalose, maltose, or sucrose (2% w/v) for 6.5 m. Trehalose administration prevented a reduction in tyrosine hydroxylase immunoreactivity in the substantia nigra (−25%), striatum (−38%), and gut (−18%) in PrP-A53T mice. It also modulated the gut microbiota, reducing the loss of diversity seen in PrP-A53T mice and promoting bacteria negatively correlated with PD in patients. Additionally, trehalose treatment increased the intestinal secretion of glucagon-like peptide 1 (GLP-1) by 29%. Maltose and sucrose, which break down into glucose, did not show neuroprotective effects, suggesting glucose is not involved in trehalose-mediated neuroprotection. Since trehalose is unlikely to cross the intestinal barrier at the given dose, the results suggest its effects are mediated indirectly through the gut microbiota and GLP-1. Full article
(This article belongs to the Section Nutrition and Metabolism)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Experimental timeline in mice showing the distribution of plain water or water mixed with 2% <span class="html-italic">w</span>/<span class="html-italic">v</span> of trehalose, maltose, or sucrose and in vivo tests performed. Stool harvest (SH); open field (OF); month (m). (<b>B</b>) Chemical formula of administered sugars: trehalose, maltose, and sucrose. Differences between the formula of trehalose and other sugars are framed in red.</p>
Full article ">Figure 2
<p>Effect of genotype and treatments on mice weight and glycemia. (<b>A</b>) Time course of mice’s weight from 3 m to 9.5 m of age. (<b>B</b>) Effect of sugar administration on blood sugar levels at the moment of euthanasia when mice are 9.5 m old. n = 7–8 mice per group. # <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 for the genotype effect (NC vs. PrP-A53T), * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01; two-way ANOVA, Tukey test.</p>
Full article ">Figure 3
<p>Effect of genotype and treatments on motor behavior. (<b>A</b>) Representation of mice trajectories during the open field test (30 min) at 9.5 m of age. (<b>B</b>) Analysis of the total distance traveled by mice in the open field test from 3 m to 9.5 m of age. (<b>C</b>) Analysis of the activity levels of mice during the open field test by measuring the percentage of active time from 3 m to 9.5 m of age. (<b>D</b>) Number of entries in the open field inner zone to measure anxiety-like behavior from 3 m to 9.5 m of age. n = 7–8 mice per group. ## <span class="html-italic">p</span> &lt; 0.01, #### <span class="html-italic">p</span> &lt; 0.0001 for the genotype effect (NC vs. PrP-A53T); two-way ANOVA, Tukey test.</p>
Full article ">Figure 4
<p>Effect of genotype and treatments on anxiety and depression. (<b>A</b>) Representative images of nest building scores. Scale bar = 1 cm. (<b>B</b>) Nest-building scores were assessed at 1, 4, 8, 12, 24, and 48 h. n = 7–8 mice per group. # <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 for the genotype effect (NC vs. PrP-A53T); two-way ANOVA, Tukey test.</p>
Full article ">Figure 5
<p>Effect of genotype and treatments on mice stool. (<b>A</b>) Fecal boli count of mice in a 30-min time period at 9.5 m of age. (<b>B</b>) Total wet mass of the feces harvested at 9.5 m of age. (<b>C</b>) Percentage of water present in the harvested feces at 9.5 m of age. n = 7–8 mice per group. ### <span class="html-italic">p</span> &lt; 0.001, #### <span class="html-italic">p</span> &lt; 0.0001 for the genotype effect (NC vs. PrP-A53T); two-way ANOVA, Tukey test.</p>
Full article ">Figure 6
<p>Effect of genotype and treatments on phosphorylated S129 human α-syn mean intensity in the substantia nigra pars compacta (SNpc) and the striatum. (<b>A</b>) Representative example of α-syn immunofluorescence in the SNpc. Scale bar = 20 µm. (<b>B</b>) Representative example of striatal α-syn immunofluorescence. Scale bar = 500 µm. (<b>C</b>) Phosphorylated S129 human α-syn mean intensity in the SNpc in PrP-A53T mice and NC. Values shown are the mean pixel intensity ± SEM of n = 7–8 mice per group. (<b>D</b>) Striatal phosphorylated S129 human α-syn mean intensity in PrP-A53T mice and NC. Values shown are the mean pixel intensity ± SEM of 7–8 mice per group. **** <span class="html-italic">p</span> &lt; 0.0001; two-way ANOVA, Tukey test.</p>
Full article ">Figure 7
<p>Effect of genotype and treatments on TH in the substantia nigra pars compacta (SNpc), the striatum and the myenteric plexus. (<b>A</b>) Representative microphotographs of TH immunofluorescence in the SNpc. Scale bar = 20 µm. (<b>B</b>) TH mean intensity in the SNpc in PrP-A53T mice and NC. Values shown are the mean pixel intensity ± SEM of n = 7–8 mice per group. (<b>C</b>) Representative microphotographs of striatal TH immunofluorescence. Scale bar = 500 µm. (<b>D</b>) Striatal TH mean intensity in PrP-A53T mice and NC. Values shown are the mean pixel intensity ± SEM of 7–8 mice per group. (<b>E</b>) Representative microphotographs of TH immunofluorescence in the myenteric plexus. Scale bar = 20 µm. (<b>F</b>) TH mean intensity in the myenteric plexus in PrP-A53T mice and NCs. Values shown are the mean pixel intensity ± SEM of 7–8 mice per group. * <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; 0.01, #### <span class="html-italic">p</span> &lt; 0.0001 for the genotype effect (NC vs. PrP-A53T); two-way ANOVA, Tukey test.</p>
Full article ">Figure 8
<p>Variation in gut microbiota composition by measuring the phylogenetic beta diversity between the various groups of mice. (<b>A</b>) PCA of Euclidian distances calculated using PhILR-transformed abundances for the NCs and PrP-A53T mice treated with water at the genus and family levels. Each point represents a unique gut microbiome sample (n = 7–8). Ellipses show the normal-theory confidence regions. Principal components one and two explained 79.9% and 11.1% of the variation in gut microbiome structure at the genus level and 73.2% and 23.2% at the family level, respectively. <span class="html-italic">p</span> &lt; 0.001 for each graph. (<b>B</b>) Beta-diversity boxplots and indices for water, trehalose, maltose, and sucrose in NC-treated mice at the genus and family levels. Tukey HSD test group letters (<span class="html-italic">p</span> &lt; 0.05) are represented: groups with the same letter are not different; groups with different letters are significantly different. (<b>C</b>) Beta-diversity boxplots and indices for water, trehalose, maltose, and sucrose in PrP-A53T-treated mice at the genus and family levels. n = 7–8 mice per group. Tukey HSD test group letters (<span class="html-italic">p</span> &lt; 0.05) are represented: groups with the same letter are not different; groups with different letters are significantly different.</p>
Full article ">Figure 9
<p>Effect of genotype and treatment on the relative abundance of taxa in NC and PrP-A53T mice using LEfSe analysis. (<b>A</b>) Relative abundance of bacteria with LDA scores of differentially abundant operational taxonomic units (OTUs) among NC mice (green) or PrP-A53T mice (brown) treated with normal water. (<b>B</b>) Relative abundance of bacteria with LDA scores of differentially abundant OTUs among NC mice (green) or PrP-A53T mice (brown) treated with trehalose. (<b>C</b>) Relative abundance of bacteria with LDA scores of differentially abundant OTUs among NC mice treated with normal water (grey), trehalose (blue), maltose (orange), and sucrose (rose). (<b>D</b>) Relative abundance of bacteria with LDA scores of differentially abundant OTUs among PrP-A53T mice treated with normal water (grey), trehalose (blue), maltose (orange), and sucrose (rose). The LDA scores represent the effect size of each abundant OTUs. Species enriched in each group with an LDA score &gt; 2 are considered. n = 7–8 mice per group.</p>
Full article ">Figure 10
<p>Effect of genotype and treatments on gut microbiota functional markers and metabolic pathways. (<b>A</b>) Sankey diagram (or alluvial plot) showing the bacterial contribution to predicted functional markers in the PrP-A53T trehalose-treated mice, from feces harvested at 9.5 m of age. Left: functional microbiota selected markers. Middle: functional class or superclass. Right: contribution of the most abundant families in the microbiota. The wider the band, the greater the contribution. The functional fraction was calculated by accumulating the genome coverage values of genomes from a specific microbial group that possesses a given functional trait. The width of curved lines from a specific microbial group to a given functional trait indicates their corresponding proportional contribution to a specific metabolism. (<b>B</b>) Fecal acetate contents at 6 m and 9.5 m of age. (<b>C</b>) Fecal propionate contents at 6 m and 9.5 m of age. (<b>D</b>) Fecal isobutyrate contents at 6 m and 9.5 m of age. n = 7–8 mice per group. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, <span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.001, <span>$</span><span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.0001 for the age effect (6 m vs. 9.5 m); three-way ANOVA, Tukey test.</p>
Full article ">Figure 11
<p>Effect of genotype and treatments on glucagon-like peptide 1 (GLP-1) immunoreactivity in the myenteric plexus. (<b>A</b>) Representative example of GLP-1 immunofluorescence in the myenteric plexus. Scale bar = 20 µm. (<b>B</b>) GLP-1 mean intensity in the myenteric plexus in PrP-A53T mice and NCs. Values shown are the mean pixel intensity ± SEM of n = 7–8 mice per group. * <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 for the genotype effect (NCs vs. PrP-A53T); two-way ANOVA, Tukey test.</p>
Full article ">Figure 12
<p>Schematic representation of potential neuroprotective pathways for trehalose in the enteric and central nervous systems in the PrP-A53T mouse model. Abbreviation: TH, tyrosine hydroxylase.</p>
Full article ">
18 pages, 968 KiB  
Review
Asymmetry in Atypical Parkinsonian Syndromes—A Review
by Patryk Chunowski, Natalia Madetko-Alster and Piotr Alster
J. Clin. Med. 2024, 13(19), 5798; https://doi.org/10.3390/jcm13195798 - 28 Sep 2024
Viewed by 358
Abstract
Background/Objectives: Atypical parkinsonian syndromes (APSs) are a group of neurodegenerative disorders that differ from idiopathic Parkinson’s disease (IPD) in their clinical presentation, underlying pathology, and response to treatment. APSs include conditions such as multiple system atrophy (MSA), progressive supranuclear palsy (PSP), corticobasal syndrome [...] Read more.
Background/Objectives: Atypical parkinsonian syndromes (APSs) are a group of neurodegenerative disorders that differ from idiopathic Parkinson’s disease (IPD) in their clinical presentation, underlying pathology, and response to treatment. APSs include conditions such as multiple system atrophy (MSA), progressive supranuclear palsy (PSP), corticobasal syndrome (CBS), and dementia with Lewy bodies (DLB). These disorders are characterized by a combination of parkinsonian features and additional symptoms, such as autonomic dysfunction, supranuclear gaze palsy, and asymmetric motor symptoms. Many hypotheses attempt to explain the causes of neurodegeneration in APSs, including interactions between environmental toxins, tau or α-synuclein pathology, oxidative stress, microglial activation, and vascular factors. While extensive research has been conducted on APSs, there is a limited understanding of the symmetry in these diseases, particularly in MSA. Neuroimaging studies have revealed metabolic, structural, and functional abnormalities that contribute to the asymmetry in APSs. The asymmetry in CBS is possibly caused by a variable reduction in striatal D2 receptor binding, as demonstrated in single-photon emission computed tomography (SPECT) examinations, which may explain the disease’s asymmetric manifestation and poor response to dopaminergic therapy. In PSP, clinical dysfunction correlates with white matter tract degeneration in the superior cerebellar peduncles and corpus callosum. MSA often involves atrophy in the pons, putamen, and cerebellum, with clinical symmetry potentially depending on the symmetry of the atrophy. The aim of this review is to present the study findings on potential symmetry as a tool for determining potential neuropsychological disturbances and properly diagnosing APSs to lessen the misdiagnosis rate. Methods: A comprehensive review of the academic literature was conducted using the medical literature available in PubMed. Appropriate studies were evaluated and examined based on patient characteristics and clinical and imaging examination outcomes in the context of potential asymmetry. Results: Among over 1000 patients whose data were collected, PSP-RS was symmetrical in approximately 84% ± 3% of cases, with S-CBD showing similar results. PSP-P was symmetrical in about 53–55% of cases, while PSP-CBS was symmetrical in fewer than half of the cases. MSA-C was symmetrical in around 40% of cases. It appears that MSA-P exhibits symmetry in about 15–35% of cases. CBS, according to the criteria, is a disease with an asymmetrical clinical presentation in 90–99% of cases. Similar results were obtained via imaging methods, but transcranial sonography produced different results. Conclusions: Determining neurodegeneration symmetry may help identify functional deficits and improve diagnostic accuracy. Patients with significant asymmetry in neurodegeneration may exhibit different neuropsychological symptoms based on their individual brain lateralization, impacting their cognitive functioning and quality of life. Full article
Show Figures

Figure 1

Figure 1
<p>Increase in the percentage of symmetry in different types of APSs. The figure is for illustrative purposes. Red—&lt;50% symmetry rate; Blue—&gt;50% symmetry rate.</p>
Full article ">Figure 2
<p>Percentage distribution of the symmetrical occurrence of clinical symptoms in the various types of APSs. CBS is presented collectively (except for S-CBS) due to the very rare occurrence of symmetrical clinical symptoms. Red—&lt;50% symmetry rate; Blue—&gt;50% symmetry rate.</p>
Full article ">
26 pages, 51990 KiB  
Article
Methamphetamine-Induced Blood Pressure Sensitization Correlates with Morphological Alterations within A1/C1 Catecholamine Neurons
by Carla Letizia Busceti, Domenico Bucci, Antonio Damato, Massimiliano De Lucia, Eleonora Venturini, Michela Ferrucci, Gloria Lazzeri, Stefano Puglisi-Allegra, Mariarosaria Scioli, Albino Carrizzo, Ferdinando Nicoletti, Carmine Vecchione and Francesco Fornai
Int. J. Mol. Sci. 2024, 25(19), 10282; https://doi.org/10.3390/ijms251910282 - 24 Sep 2024
Viewed by 320
Abstract
Methamphetamine (METH) is a drug of abuse, which induces behavioral sensitization following repeated doses. Since METH alters blood pressure, in the present study we assessed whether systolic and diastolic blood pressure (SBP and DBP, respectively) are sensitized as well. In this context, we [...] Read more.
Methamphetamine (METH) is a drug of abuse, which induces behavioral sensitization following repeated doses. Since METH alters blood pressure, in the present study we assessed whether systolic and diastolic blood pressure (SBP and DBP, respectively) are sensitized as well. In this context, we investigated whether alterations develop within A1/C1 neurons in the vasomotor center. C57Bl/6J male mice were administered METH (5 mg/kg, daily for 5 consecutive days). Blood pressure was measured by tail-cuff plethysmography. We found a sensitized response both to SBP and DBP, along with a significant decrease of catecholamine neurons within A1/C1 (both in the rostral and caudal ventrolateral medulla), while no changes were detected in glutamic acid decarboxylase. The decrease of catecholamine neurons was neither associated with the appearance of degeneration-related marker Fluoro-Jade B nor with altered expression of α-synuclein. Rather, it was associated with reduced free radicals and phospho-cJun and increased heat shock protein-70 and p62/sequestosome within A1/C1 cells. Blood pressure sensitization was not associated with altered arterial reactivity. These data indicate that reiterated METH administration may increase blood pressure persistently and may predispose to an increased cardiovascular response to METH. These data may be relevant to explain cardiovascular events following METH administration and stressful conditions. Full article
(This article belongs to the Special Issue Molecular Studies of Mutations Related to Neurodegenerative Diseases)
Show Figures

Figure 1

Figure 1
<p>Methamphetamine (METH) induces SBP and DBP sensitization in mice. (<b>A</b>) Diagram showing the protocol used for monitoring SBP and DBP values in mice subjected to daily intraperitoneal (i.p.) injection of saline or METH (5 mg/kg) for 5 days and then re-challenged with saline or METH following 6 days of withdrawal on day 11. SBP and DBP were monitored by the tail-cuff plethysmography before treatments (Basal) and 30 min after each injection with saline or METH. SBP and DBP values monitored in mice treated with saline or METH for 5 days are shown in (<b>B</b>,<b>B’</b>)<b>,</b> respectively. Values are the means ± S.E.M. <span class="html-italic">p</span> &lt; 0.05, (*) vs. saline; (#) vs. Basal, day 1 and day 2 (two-way ANOVA for repeated measures with Bonferroni’s test). Difference in SBP values (Δ SBP) and DBP values (Δ DBP) monitored at 30 min after saline or METH injections between days 5 and day 1 (Day 5–Day 1) is shown in (<b>C</b>,<b>C’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (unpaired two-tailed Student’s <span class="html-italic">t</span>-test). SBP and DBP values monitored in mice at 30 min after the challenge with saline or METH performed following 6 days of withdrawal on day 11 are shown in (<b>D</b>,<b>D’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (unpaired two-tailed Student’s <span class="html-italic">t</span>-test). Difference in SBP values (Δ SBP) and DBP values (Δ DBP) monitored at 30 min after saline or METH injections between days 11 and day 1 (Day 11–Day 1) are shown in (<b>E</b>,<b>E’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (unpaired two-tailed Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>Reiterated exposure to METH in mice does not lead to mesenteric arteries’ endothelial disfunction. Mesenteric arteries were isolated from mice subjected to repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and then sacrificed 60 min after the challenge, carried out on day 11. Phenylephrine-induced vasoconstriction and acetylcholine (ACh)-induced vasorelaxation in mesenteric arteries isolated from saline or METH-treated mice are shown in (<b>A</b>,<b>B</b>), respectively. Values are means ± S.E.M. (<b>C</b>) Double fluorescent staining for the catecholaminergic marker tyrosine hydroxylase (TH) and the endothelial marker von Willebrand factor (vWF) in mesenteric arteries isolated from mice treated with saline or METH and sacrificed 60 min following the challenge with saline or METH performed following 6 days of withdrawal on day 11. Densitometric quantification of the vWF or TH immunoreactivity (ir) is shown in (<b>D</b>,<b>E</b>), respectively. Values of optical density (OD) are means ± S.E.M.</p>
Full article ">Figure 3
<p>Within the RVLM, repeated METH administration reduces TH immunopositive cells without affecting GAD immunopositive cells. Immunohistochemical analysis for the catecholamine marker TH was carried out from dissected brains of mice subjected to repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and then sacrificed 60 min after the METH/saline challenge, which was carried out on day 11. Representative images of TH-positive neurons within A1/C1 in the RVLM of mice treated with saline or METH are shown in (<b>A</b>). The graph in (<b>B</b>) shows TH-positive cell density (n/mm<sup>2</sup>) counted in saline- and METH-treated mice. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (unpaired two-tailed Student’s <span class="html-italic">t</span>-test). The correlation analyses between TH-positive cell density within A1/C1 in the RVLM and SBP are shown in (<b>C</b>,<b>D</b>), respectively. These correlation analyses were carried out considering SBP values that were measured at 30 min following the fifth saline/METH injection or following the METH/saline challenge. * <span class="html-italic">p</span> &lt; 0.05, Pearson correlation test. (<b>E</b>) Representative images of double immunostaining for TH and GAD in the RVLM of mice following repeated injections of saline or METH (5 mg/kg, i.p., for 5 days). These mice were sacrificed 60 min following the METH-/saline challenge, carried out on day 11. The densitometry (OD: Optical Density) of GAD65/67 immunoreactivity (ir) is shown in (<b>F</b>). Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann-Whitney test).</p>
Full article ">Figure 4
<p>Within CVLM, repeated METH administration reduces TH immunopositive cells without affecting GAD immunopositive cells. Immunohistochemical analysis for TH was performed in dissected brains of mice subjected to repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and then sacrificed 60 min after the challenge, carried out on day 11. Representative images or TH-positive neurons in A1/C1 within the CVLM of mice treated with saline or METH are shown in (<b>A</b>)<b>.</b> The graph in (<b>B</b>) shows the TH-positive cell density (n/mm<sup>2</sup>) counted in saline- and METH-treated mice. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (unpaired two-tailed Student’s <span class="html-italic">t</span>-test). The correlation analysis between values of TH-positive cell density in A1/C1 of the CVLM and the respective SBP values monitored 30 min after the fifth repeated saline/METH injection or following the challenge carried out on day 11 is shown in (<b>C</b>,<b>D</b>), respectively. * <span class="html-italic">p</span> &lt; 0.05, Pearson correlation test. (<b>E</b>) Representative images of double immunostaining for TH and GAD in the CVLM of mice following repeated injections of saline or METH (5 mg/kg, i.p., for 5 days). These mice were sacrificed 60 min following the METH/saline challenge, carried out on day 11. The densitometry (OD: Optical Density) of GAD65/67 immunoreactivity (ir) is shown in (<b>F</b>). Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann-Whitney test).</p>
Full article ">Figure 5
<p>Within A2/C2 and AP, METH does not decrease catecholamine neurons. Immunohistochemical analysis for TH was performed in dissected brains of mice following repeated METH administration (5 mg/kg, i.p., for 5 days). Mice were sacrificed 60 min following the METH challenge, which was carried out on day 11. TH-positive cell density (n/mm<sup>2</sup>) counted in saline- and METH-treated mice is shown in (<b>A</b>,<b>B</b>), respectively. Values are the means ± S.E.M. Representative images or TH-positive neurons in A2/C2 and AP of mice treated with saline or METH are shown in (<b>C</b>).</p>
Full article ">Figure 6
<p>Lack of neuronal loss in the RVLM and CVLM of mice treated with METH. (<b>A</b>) Fluoro-Jade B staining was carried out within the RVLM and CVLM of mice treated with METH and sacrificed at the following different time points: (i) naïve mice (Basal); (ii) 24 h after a single treatment with METH (5 mg/kg, i.p.) (day 2); (iii) 24 h after the last injection in mice subjected to repeated injection with METH (5 mg/kg, i.p., for 3 days) (day 4); (iv) 60 min the challenge with METH (5 mg/kg, i.p.) performed following 6 days of withdrawal after a repeated treatment with METH (5 mg/kg, i.p., for 5 days) (day 11). (<b>B</b>) Immunohistochemical analysis for the glial fibrillary acidic protein (GFAP) in the RVLM and CVLM of mice subjected to a single or repeated injection of METH and sacrificed at different time intervals (see above).</p>
Full article ">Figure 7
<p>METH does not increase αSyn within A1/C1 neurons of the RVLM and CVLM. Double immunofluorescence for the catecholamine marker TH and αSyn was performed in the RVLM and CVLM of mice following repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and sacrificed 60 min after the METH challenge, which was carried out on day 11. Representative images of the double immunostaining for TH and αSyn in A1/C1 within the RVLM and CVLM of mice treated with saline or METH are shown in (<b>A</b>,<b>B</b>), respectively.</p>
Full article ">Figure 8
<p>Repeated administrations of METH lead to intracellular reduction of free radicals in catecholamine A1/C1 neurons of the RVLM and CVLM. Double fluorescent analysis for TH-immunoreactivity for the catecholamine marker TH and αSyn was performed in the RVLM and CVLM of mice subjected to repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and sacrificed 60 min after the challenge carried out, on day 11. Representative images of the double immunostaining for TH and αSyn in A1/C1 within the RVLM and CVLM of mice treated with saline or METH are shown in (<b>A</b>,<b>B</b>), respectively. The respective densitometric values of DHE stain optical density (OD) are shown in (<b>A’</b>,<b>B’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann–Whitney test).</p>
Full article ">Figure 9
<p>Repeated administrations of METH increase HSP70 within A1/C1 neurons. Representative images of the double immunostaining for TH and HSP70 in A1/C1 within the RVLM and CVLM of mice treated with saline or METH are shown in (<b>A</b>,<b>B</b>), respectively. The densitometry values (OD: Optical Density) of HSP70 immunoreactivity (ir) are shown in (<b>A’</b>,<b>B’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann–Whitney test).</p>
Full article ">Figure 10
<p>Following METH administration, p62 increases within A1/C1 neurons of the RVLM and CVLM. Double immunofluorescent analysis for TH and p62 was performed in the RVLM and CVLM of mice some days following METH administration. Representative images of double immunostaining for TH and p62 in A1/C1 within the RVLM and CVLM of mice treated with saline or METH are shown in (<b>A</b>,<b>B</b>), respectively. The values of p62 are shown in (<b>A’</b>,<b>B’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann–Whitney test).</p>
Full article ">Figure 11
<p>Repeated administrations of METH lead to intracellular increased p-cJun expression in catecholamine A1/C1 neurons of the RVLM and CVLM. Double immunofluorescent analysis for TH and p-cJun was performed in the RVLM and CVLM of mice subjected to repeated injections of saline or METH (5 mg/kg, i.p., for 5 days) and sacrificed 60 min after the challenge, carried out on day 11. Representative images of the double immunostaining for TH and p-cJun in A1/C1 within the RVLM and CVLM of mice treated with saline or METH are shown in (<b>A</b>,<b>B</b>), respectively. The respective densitometric values (OD: Optical Density) of p-cJun intracellular immunoreactivity (ir) are shown in (<b>A’</b>,<b>B’</b>), respectively. Values are the means ± S.E.M. * <span class="html-italic">p</span> &lt; 0.05 (two-tailed Mann–Whitney test).</p>
Full article ">Figure 12
<p>Diagram showing a possible molecular scenario occurring in response to repeated METH exposure in A1/C1 catecholamine neurons. Our findings demonstrate that, following repeated METH exposure, increased expression of HSP70 and p62 occurs in A1/C1 catecholamine neurons in the RVLM and CVLM. The consequent antioxidant effect may be responsible for the observed reduced intracellular levels of free radicals in the same cells. This may underlie a reduced phosphorylation of the transcriptional factor cJun and a downregulated expression of TH with a consequent phenotypic shift of the A1/C1 catecholamine cells within the RVLM and CVLM. The expression of p-cJun, which is also a marker of neuronal activity, is suppressed in the withdrawal phase following METH administration. This remains a quite general marker, which should be implemented by dedicated electrophysiological studies to be carried out in specific future research projects. Dotted arrows indicate that a causal relationship between each event is not demonstrated. This hypothesis warrants further investigation.</p>
Full article ">
25 pages, 2774 KiB  
Review
Exploring the Role of Reactive Oxygen Species in the Pathogenesis and Pathophysiology of Alzheimer’s and Parkinson’s Disease and the Efficacy of Antioxidant Treatment
by Talin Gogna, Benjamin E. Housden and Annwyne Houldsworth
Antioxidants 2024, 13(9), 1138; https://doi.org/10.3390/antiox13091138 - 20 Sep 2024
Viewed by 1181
Abstract
Alzheimer’s (AD) and Parkinson’s Disease (PD) are life-altering diseases that are characterised by progressive memory loss and motor dysfunction. The prevalence of AD and PD is predicted to continuously increase. Symptoms of AD and PD are primarily mediated by progressive neuron death and [...] Read more.
Alzheimer’s (AD) and Parkinson’s Disease (PD) are life-altering diseases that are characterised by progressive memory loss and motor dysfunction. The prevalence of AD and PD is predicted to continuously increase. Symptoms of AD and PD are primarily mediated by progressive neuron death and dysfunction in the hippocampus and substantia nigra. Central features that drive neurodegeneration are caspase activation, DNA fragmentation, lipid peroxidation, protein carbonylation, amyloid-β, and/or α-synuclein formation. Reactive oxygen species (ROS) increase these central features. Currently, there are limited therapeutic options targeting these mechanisms. Antioxidants reduce ROS levels by the induction of antioxidant proteins and direct neutralisation of ROS. This review aims to assess the effectiveness of antioxidants in reducing ROS and neurodegeneration. Antioxidants enhance major endogenous defences against ROS including superoxide dismutase, catalase, and glutathione. Direct neutralisation of ROS by antioxidants protects against ROS-induced cytotoxicity. The combination of Indirect and direct protective mechanisms prevents ROS-induced α-synuclein and/or amyloid-β formation. Antioxidants ameliorate ROS-mediated oxidative stress and subsequent deleterious downstream effects that promote apoptosis. As a result, downstream harmful events including neuron death, dysfunction, and protein aggregation are decreased. The protective effects of antioxidants in human models have yet to directly replicate the success seen in cell and animal models. However, the lack of diversity in antioxidants for clinical trials prevents a definitive answer if antioxidants are protective. Taken together, antioxidant treatment is a promising avenue in neurodegenerative disease therapy and subsequent clinical trials are needed to provide a definitive answer on the protective effects of antioxidants. No current treatment strategies have significant impact in treating advanced AD and PD, but new mimetics of endogenous mitochondrial antioxidant enzymes (Avasopasem Manganese, GC4419 AVA) may be a promising innovative option for decelerating neurodegenerative progress in the future at the mitochondrial level of OS. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Diagram depicting the cytotoxic mechanism of ROS and cell defences against ROS. .O2 generated by aerobic respiration in which SOD1 converts .O<sub>2</sub> in the mitochondrial matrix whilst SOD2 converts .O in the intermitochondrial space to H O . Then catalase breaks down H O into water and oxygen or H O oxidise glutathione into glutathione disulfide in the presence of 22 glutathione reductase. Cu<sup>2+</sup>/Fe<sup>2+</sup> convert H<sub>2</sub>O<sub>2</sub> into .OH, which oxidises DNA to cause double-strand or single-strand breaks. Thioredoxins and peroxiredoxins also maintain intracellular redox homeostasis reducing lipid peroxidation and the maintenance of cell membranes. Lipid peroxidation is important in the α-synuclein pathology associated with PD. .OH also activates caspases which break apart proteins that maintain the cell cytoskeleton, DNA repair, and ATP production as well as causing lipid peroxidation and protein carbonylation which causes cell stress. Lastly, .O<sub>2</sub> damages mitochondrial DNA and causes the release of cytochrome C which activates caspases. Text in red indicates Abbreviations; .O : Superoxide anions, SOD: Superoxide dismutase, H<sub>2</sub>O<sub>2</sub> : Hydrogen peroxide, .OH: Hydroxyl radical, GSH: Glutathione, GR: Glutathione reductase, GSSG: Glutathione disulfide, and Aβ: Amyloid-β.(9,10).</p>
Full article ">Figure 2
<p>Flow chart depicting the search, screening, inclusion, and exclusion criteria in conjunction with the order followed. Papers collated from this method are referenced in the text were analysed, and relevant information was extracted and referenced where appropriate.</p>
Full article ">Figure 3
<p>Cellular protective mechanisms of key antioxidant treatments. Curcumin, VA, TFA, PCA, and 17β-oestradiol have direct experimental evidence that they activate SIRT1. As a result, SIRT1 activates transcription factors by removal of acetyl groups: PGC1α, Nrf2, and FoxO3a. PGC1α and Nrf2 increase the expression of SOD1/2 whilst FoxO3a increases catalase expression. Other antioxidants quercetin and carotenoids increase antioxidant defenses they may act through SIRT1 activation although unclear, furthermore, they both increase glutathione expression. Metal chelators prevent the generation of .OH by binding to free Cu<sup>2+</sup> and Fe<sup>2+</sup> and preventing them from participating in REDOX reactions. Abbreviations: .O<sub>2</sub>: Superoxide anions, SOD: Superoxide dismutase, H<sub>2</sub>O<sub>2</sub>: Hydrogen peroxide, .OH: Hydroxyl radical, GSH: Glutathione, GR: Glutathione reductase, GSSG: Glutathione disulfide, SIRT1: Silent information regulator 1, PGC1α: Peroxisome proliferator-activated receptor gamma coactivator.</p>
Full article ">
24 pages, 1289 KiB  
Review
Navigating the Neurobiology of Parkinson’s: The Impact and Potential of α-Synuclein
by Erlandas Paulėkas, Tadas Vanagas, Saulius Lagunavičius, Evelina Pajėdienė, Kęstutis Petrikonis and Daiva Rastenytė
Biomedicines 2024, 12(9), 2121; https://doi.org/10.3390/biomedicines12092121 - 18 Sep 2024
Viewed by 2790
Abstract
Parkinson’s disease (PD) is the second most prevalent neurodegenerative disease worldwide; therefore, since its initial description, significant progress has been made, yet a mystery remains regarding its pathogenesis and elusive root cause. The widespread distribution of pathological α-synuclein (αSyn) aggregates throughout the body [...] Read more.
Parkinson’s disease (PD) is the second most prevalent neurodegenerative disease worldwide; therefore, since its initial description, significant progress has been made, yet a mystery remains regarding its pathogenesis and elusive root cause. The widespread distribution of pathological α-synuclein (αSyn) aggregates throughout the body raises inquiries regarding the etiology, which has prompted several hypotheses, with the most prominent one being αSyn-associated proteinopathy. The identification of αSyn protein within Lewy bodies, coupled with genetic evidence linking αSyn locus duplication, triplication, as well as point mutations to familial Parkinson’s disease, has underscored the significance of αSyn in initiating and propagating Lewy body pathology throughout the brain. In monogenic and sporadic PD, the presence of early inflammation and synaptic dysfunction leads to αSyn aggregation and neuronal death through mitochondrial, lysosomal, and endosomal functional impairment. However, much remains to be understood about αSyn pathogenesis, which is heavily grounded in biomarkers and treatment strategies. In this review, we provide emerging new evidence on the current knowledge about αSyn’s pathophysiological impact on PD, and its presumable role as a specific disease biomarker or main target of disease-modifying therapies, highlighting that this understanding today offers the best potential of disease-modifying therapy in the near future. Full article
Show Figures

Figure 1

Figure 1
<p>Molecular mechanisms contributing to Parkinson’s disease. Genetic mutations, mitochondrial dysfunction, and other significant biological components negatively influence physiological condition and lead to misfolding and aggregation of α-synuclein, which contributes to formation of aggregates found in the brain, skin, gut, and other organs.</p>
Full article ">Figure 2
<p>Overview of the proposed functions of physiological and pathological forms of alpha-synuclein. VAMP2—vesicle associated membrane protein 2, SNARE—soluble N-ethylmaleimide sensitive factor, NAC—non-Aβ-amyloid component, Ser—serine, AA—amino acids, α-syn—alpha-synuclein, DA—dopamine, mt—mitochondria, ETC—electron transport chain, NMDA—N-methyl-D-aspartate, NLRP3—nucleotide-binding oligomerization domain-leucine-rich repeat-pyrin domain-containing 3.</p>
Full article ">
26 pages, 1328 KiB  
Review
From Brain to Muscle: The Role of Muscle Tissue in Neurodegenerative Disorders
by Elisa Duranti and Chiara Villa
Biology 2024, 13(9), 719; https://doi.org/10.3390/biology13090719 - 12 Sep 2024
Viewed by 1157
Abstract
Neurodegenerative diseases (NDs), like amyotrophic lateral sclerosis (ALS), Alzheimer’s disease (AD), and Parkinson’s disease (PD), primarily affect the central nervous system, leading to progressive neuronal loss and motor and cognitive dysfunction. However, recent studies have revealed that muscle tissue also plays a significant [...] Read more.
Neurodegenerative diseases (NDs), like amyotrophic lateral sclerosis (ALS), Alzheimer’s disease (AD), and Parkinson’s disease (PD), primarily affect the central nervous system, leading to progressive neuronal loss and motor and cognitive dysfunction. However, recent studies have revealed that muscle tissue also plays a significant role in these diseases. ALS is characterized by severe muscle wasting as a result of motor neuron degeneration, as well as alterations in gene expression, protein aggregation, and oxidative stress. Muscle atrophy and mitochondrial dysfunction are also observed in AD, which may exacerbate cognitive decline due to systemic metabolic dysregulation. PD patients exhibit muscle fiber atrophy, altered muscle composition, and α-synuclein aggregation within muscle cells, contributing to motor symptoms and disease progression. Systemic inflammation and impaired protein degradation pathways are common among these disorders, highlighting muscle tissue as a key player in disease progression. Understanding these muscle-related changes offers potential therapeutic avenues, such as targeting mitochondrial function, reducing inflammation, and promoting muscle regeneration with exercise and pharmacological interventions. This review emphasizes the importance of considering an integrative approach to neurodegenerative disease research, considering both central and peripheral pathological mechanisms, in order to develop more effective treatments and improve patient outcomes. Full article
(This article belongs to the Special Issue Repair and Regeneration of Skeletal Muscle)
Show Figures

Figure 1

Figure 1
<p>A schematic representation of skeletal muscle structure. The image was created with the use of Servier Medical Art modified templates, licensed under a Creative Common Attribution 3.0 Unported License (<a href="https://smart.servier.com" target="_blank">https://smart.servier.com</a>, accessed on 28 August 2024).</p>
Full article ">Figure 2
<p>A schematic representation of the accumulation of ROS leading to mitochondrial dysfunction and cellular damage. This cascade of events results in muscle atrophy and weakness. The image was created with the use of Servier Medical Art modified templates, licensed under a Creative Common Attribution 3.0 Unported License (<a href="https://smart.servier.com" target="_blank">https://smart.servier.com</a>, accessed on 28 August 2024).</p>
Full article ">Figure 3
<p>The aggregates of α-synuclein caused by alterations in protein degradation pathways promote the muscle cell alterations typically found in PD patients. The image was created with the use of Servier Medical Art modified templates, licensed under a Creative Common Attribution 3.0 Unported License (<a href="https://smart.servier.com" target="_blank">https://smart.servier.com</a>, accessed on 28 August 2024).</p>
Full article ">
15 pages, 1096 KiB  
Review
Immunotherapy for Parkinson’s Disease and Alzheimer’s Disease: A Promising Disease-Modifying Therapy
by Anns Mahboob, Hasan Ali, AlJazi AlNaimi, Mahmoud Yousef, Mlaak Rob, Nawaf Ahmad Al-Muhannadi, Degiri Kalana Lasanga Senevirathne and Ali Chaari
Cells 2024, 13(18), 1527; https://doi.org/10.3390/cells13181527 - 12 Sep 2024
Viewed by 1057
Abstract
Alzheimer’s disease (AD) and Parkinson’s disease (PD) are two neurodegenerative diseases posing a significant disease burden due to their increasing prevalence and socio-economic cost. Traditional therapeutic approaches for these diseases exist but provide limited symptomatic relief without addressing the underlying pathologies. This review [...] Read more.
Alzheimer’s disease (AD) and Parkinson’s disease (PD) are two neurodegenerative diseases posing a significant disease burden due to their increasing prevalence and socio-economic cost. Traditional therapeutic approaches for these diseases exist but provide limited symptomatic relief without addressing the underlying pathologies. This review examines the potential of immunotherapy, specifically monoclonal antibodies (mAbs), as disease-modifying treatments for AD and PD. We analyze the pathological mechanisms of AD and PD, focusing on the roles of amyloid-beta (Aβ), tau (τ), and alpha-synuclein (α-syn) proteins. We discuss the latest advancements in mAb therapies targeting these proteins, evaluating their efficacy in clinical trials and preclinical studies. We also explore the challenges faced in translating these therapies from bench to bedside, including issues related to safety, specificity, and clinical trial design. Additionally, we highlight future directions for research, emphasizing the need for combination therapies, improved biomarkers, and personalized treatment strategies. This review aims to provide insights into the current state and future potential of antibody-based immunotherapy in modifying the course of AD and PD, ultimately improving patient outcomes and quality of life. Full article
Show Figures

Figure 1

Figure 1
<p>Historical timeline of AD.</p>
Full article ">Figure 2
<p>A representation of the multifaceted pathophysiology of Parkinson’s disease (PD).</p>
Full article ">
15 pages, 1354 KiB  
Article
Postencephalitic Parkinsonism: Unique Pathological and Clinical Features—Preliminary Data
by Sabrina Strobel, Jeswinder Sian-Hulsmann, Dennis Tappe, Kurt Jellinger, Peter Riederer and Camelia-Maria Monoranu
Cells 2024, 13(18), 1511; https://doi.org/10.3390/cells13181511 - 10 Sep 2024
Viewed by 552
Abstract
Postencephalitic parkinsonism (PEP) is suggested to show a virus-induced pathology, which is different from classical idiopathic Parkinson’s disease (PD) as there is no α-synuclein/Lewy body pathology. However, PEP shows a typical clinical representation of motor disturbances. In addition, compared to PD, there is [...] Read more.
Postencephalitic parkinsonism (PEP) is suggested to show a virus-induced pathology, which is different from classical idiopathic Parkinson’s disease (PD) as there is no α-synuclein/Lewy body pathology. However, PEP shows a typical clinical representation of motor disturbances. In addition, compared to PD, there is no iron-induced pathology. The aim of this preliminary study was to compare PEP with PD regarding iron-induced pathology, using histochemistry methods on paraffin-embedded post-mortem brain tissue. In the PEP group, iron was not seen, except for one case with sparse perivascular depositions. Rather, PEP offers a pathology related to tau-protein/neurofibrillary tangles, with mild to moderate memory deficits only. It is assumed that this virus-induced pathology is due to immunological dysfunctions causing (neuro)inflammation-induced neuronal network disturbances as events that trigger clinical parkinsonism. The absence of iron deposits implies that PEP cannot be treated with iron chelators. The therapy with L-Dopa is also not an option, as L-Dopa only leads to an initial slight improvement in symptoms in isolated cases. Full article
Show Figures

Figure 1

Figure 1
<p>Iron depositions: (<b>a</b>) perivascular (arrows); (<b>b</b>) intraneuronal (arrow); (<b>c</b>) microglia (white arrow), oligodendrocyte (green arrow); (<b>d</b>) sparsely perivascular deposition; (<b>a</b>–<b>c</b>) PD cases; d PEP case (magnification 100×).</p>
Full article ">Figure 2
<p>Pathophysiological pathways in Parkinson’s disease and postencephalitic parkinsonism.</p>
Full article ">
18 pages, 10927 KiB  
Article
Transient Increases in Neural Oscillations and Motor Deficits in a Mouse Model of Parkinson’s Disease
by Yue Wu, Lidi Lu, Tao Qing, Suxin Shi and Guangzhan Fang
Int. J. Mol. Sci. 2024, 25(17), 9545; https://doi.org/10.3390/ijms25179545 - 2 Sep 2024
Viewed by 674
Abstract
Parkinson’s disease (PD) is a neurodegenerative disorder characterized by motor symptoms like tremors and bradykinesia. PD’s pathology involves the aggregation of α-synuclein and loss of dopaminergic neurons, leading to altered neural oscillations in the cortico-basal ganglia-thalamic network. Despite extensive research, the relationship between [...] Read more.
Parkinson’s disease (PD) is a neurodegenerative disorder characterized by motor symptoms like tremors and bradykinesia. PD’s pathology involves the aggregation of α-synuclein and loss of dopaminergic neurons, leading to altered neural oscillations in the cortico-basal ganglia-thalamic network. Despite extensive research, the relationship between the motor symptoms of PD and transient changes in brain oscillations before and after motor tasks in different brain regions remain unclear. This study aimed to investigate neural oscillations in both healthy and PD model mice using local field potential (LFP) recordings from multiple brain regions during rest and locomotion. The histological evaluation confirmed the significant dopaminergic neuron loss in the injection side in 6-OHDA lesioned mice. Behavioral tests showed motor deficits in these mice, including impaired coordination and increased forelimb asymmetry. The LFP analysis revealed increased delta, theta, alpha, beta, and gamma band activity in 6-OHDA lesioned mice during movement, with significant increases in multiple brain regions, including the primary motor cortex (M1), caudate–putamen (CPu), subthalamic nucleus (STN), substantia nigra pars compacta (SNc), and pedunculopontine nucleus (PPN). Taken together, these results show that the motor symptoms of PD are accompanied by significant transient increases in brain oscillations, especially in the gamma band. This study provides potential biomarkers for early diagnosis and therapeutic evaluation by elucidating the relationship between specific neural oscillations and motor deficits in PD. Full article
Show Figures

Figure 1

Figure 1
<p>Dopaminergic cell loss in the SNc after 6-OHDA injection. (<b>A</b>) Schematic representation of injection sites in the SNc, and the subplots C-H show enlarged pictures of the boxed area. (<b>B</b>) The number of TH+ neurons in the SNc for the sham group (n = 10) and the lesion group (n = 10). (<b>C</b>–<b>H</b>) Immunofluorescence staining of TH+ neurons (green, <b>C</b>,<b>F</b>), DAPI (blue, <b>D</b>,<b>G</b>), and a merged image (<b>E</b>,<b>H</b>) of the injection side of the SNc. The white arrows indicate TH+ neurons in the pictures. The scale bar represents 100 μm. All data are expressed as the mean ± SD. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Behavioral performance of mice after 6-OHDA injection. (<b>A</b>–<b>C</b>) Latency to fall, speed at fall, and total distance traveled in the rotarod test. (<b>D</b>–<b>E</b>) Total time for traversal and number of hindlimb lapses for the balance beam traversal test. (<b>F</b>) Time to descend from top to base in the pole test. n = 16 for the lesion group and n = 15 for the sham group. All data are expressed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Limb symmetry, gait impairment, and spontaneous locomotion after 6-OHDA injection. (<b>A</b>) Percentage of forelimb use on the injection side in the rearing cylinder test. n = 16 for the lesion group and n = 12 for the sham group. (<b>B</b>,<b>C</b>) Percentage of diagonal support and three-limb support during locomotion. (<b>D</b>–<b>F</b>) Average distance moved, number of entries into the central area, and percentage of time spent in the central area in the open field test. For pictures (<b>B</b>–<b>F</b>), n = 16 for the lesion group and n = 15 for the sham group. All data are expressed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>Verification of electrode placement. (<b>A</b>) Primary motor cortex (M1). (<b>B</b>) Corpus striatum (CPu). (<b>C</b>) Substantia nigra pars compacta (SNc). (<b>D</b>) Subthalamic nucleus (STN). (<b>E</b>) Pedunculopontine tegmental nucleus (PPN). Arrows indicate electrode sites. The scale bar represents 500 μm.</p>
Full article ">Figure 5
<p>Typical LFP waveforms recorded from the M1, CPu, STN, SNc, and PPN during resting and walking states for both sham and lesion mice.</p>
Full article ">Figure 6
<p>Power spectral density of LFPs recorded for the M1 (<b>A</b>), CPu (<b>B</b>), STN (<b>C</b>), PPN (<b>D</b>), and SNc (<b>E</b>). Left column shows average power spectral density of sham group during resting (red) and walking (blue) states for each brain region; middle column indicates average power spectral density of lesion group during resting (red) and walking (blue) states for each brain region; right column illustrates differences in power spectral density between walking and resting states in sham group (green) and lesion group (orange). All data are expressed as mean ± SEM. The dotted lines divide the frequency range into the delta (1–4 Hz), theta (4–8 Hz), alpha (8–21 Hz), beta (21–32 Hz), and gamma (32–100 Hz) bands.</p>
Full article ">Figure 7
<p>Difference in power spectra of various bands of LFP, including delta (1–4 Hz) (<b>A</b>), theta (4–8 Hz) (<b>B</b>), alpha (8–21 Hz) (<b>C</b>), beta (21–32 Hz) (<b>D</b>), and gamma (32–100 Hz) (<b>E</b>) bands. n = 10 for each group. All data are expressed as the mean ± SD. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
14 pages, 3209 KiB  
Article
Therapeutic Effect of Padina arborescens Extract on a Cell System Model for Parkinson’s Disease
by Dong Hwan Ho, Hyejung Kim, Daleum Nam, Mi Kyoung Seo, Sung Woo Park, Dong-Kyu Kim and Ilhong Son
NeuroSci 2024, 5(3), 301-314; https://doi.org/10.3390/neurosci5030024 - 30 Aug 2024
Viewed by 623
Abstract
Leucine-rich repeat kinase 2 (LRRK2) and α-synuclein are involved in the pathogenesis of Parkinson’s disease. The activity of LRRK2 in microglial cells is associated with neuroinflammation, and LRRK2 inhibitors are crucial for alleviating this neuroinflammatory response. α-synuclein contributes to oxidative stress in the [...] Read more.
Leucine-rich repeat kinase 2 (LRRK2) and α-synuclein are involved in the pathogenesis of Parkinson’s disease. The activity of LRRK2 in microglial cells is associated with neuroinflammation, and LRRK2 inhibitors are crucial for alleviating this neuroinflammatory response. α-synuclein contributes to oxidative stress in the dopaminergic neuron and neuroinflammation through Toll-like receptors in microglia. In this study, we investigated the effect of the marine alga Padina arborescens on neuroinflammation by examining LRRK2 activation and the aggregation of α-synuclein. P. arborescens extract inhibits LRRK2 activity in vitro and decreases lipopolysaccharide (LPS)-induced LRRK2 upregulation in BV2, a mouse microglial cell line. Treatment with P. arborescens extract decreased tumor necrosis factor-α (TNF-α) gene expression by LPS through LRRK2 inhibition in BV2. It also attenuated TNF-α gene expression, inducible nitric oxide synthase, and the release of TNF-α and cellular nitric oxide in rat primary microglia. Furthermore, P. arborescens extract prevented rotenone (RTN)-induced oxidative stress in primary rat astrocytes and inhibited α-synuclein fibrilization in an in vitro assay using recombinant α-synuclein and in the differentiated human dopaminergic neuronal cell line SH-SY5Y (dSH). The extract increased lysosomal activity in dSH cells. In addition, P. arborescens extract slightly prolonged the lifespan of Caenorhabditis elegans, which was reduced by RTN treatment. Full article
Show Figures

Figure 1

Figure 1
<p>PA_EXT inhibits LRRK2 activity. (<b>A</b>) PA_EXT and a commercially available LRRK2 inhibitor, MLi-2, were tested for their effects on autophosphorylation at the serine 1292 site (pS1292) of the LRRK2 G2019S recombinant protein. Densitometry reading of pS1292 was normalized to total LRRK2 densitometry reading, and each test group was compared to the control (DMSO). (<b>B</b>) S1292 autophosphorylation levels were analyzed and graphed. One-way analysis of variance (ANOVA) with Bonferroni’s multiple comparison test was used, <span class="html-italic">n</span> = 2, *; <span class="html-italic">p</span> &lt; 0.05. n.s; not significant. (<b>C</b>) LRRK2 activity changes after treatment with 200 ng/mL LPS or 1 μg/mL PA_EXT for 18 h in BV2 cells were assessed by the levels of phosphorylation at the serine 935 site (pS935) in mice using Western blot analysis. (<b>D</b>) S935 phosphorylation levels were analyzed and graphed. The densitometry reading of pS935 was normalized to the total LRRK2 densitometry reading, and each group was compared to the controls (DMSO and vehicle). A two-way ANOVA with Bonferroni’s multiple comparison test was used, n = 3, *; <span class="html-italic">p</span> &lt; 0.05, ****; <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>PA_EXT reduces pro-inflammatory cytokines in BV2 cells. (<b>A</b>) TNF-α gene expression levels after treatment with 200 ng/mL LPS or 1 μg/mL PA_EXT for 4 h were measured using qPCR and compared to the controls (DMSO and vehicle). (<b>B</b>) The phosphorylation of LRRK2 in BV2 was analyzed using Western blot analysis. (<b>C</b>) Graph of LRRK2 phosphorylation data. The densitometry reading of pS935 was normalized to that of total LRRK2 and compared to the controls (DMSO and vehicle). One-way ANOVA with Bonferroni’s multiple comparison test was used, <span class="html-italic">n</span> = 3, *; <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, n.s; not significant.</p>
Full article ">Figure 3
<p>PA_EXT reduces pro-inflammatory cytokines in rat primary microglia. Rat primary microglia cells were treated with 100 ng/mL LPS or 0.8 μg/mL PA_EXT. Gene expression levels of TNF-α (<b>A</b>) and inducible iNOS (<b>B</b>) were analyzed using qPCR. (<b>C</b>) Supernatants of cell lysates were analyzed using a commercial rat TNF-α ELISA kit. (<b>D</b>) NO levels were measured using Greiss assay (<b>D</b>) for nitrite (NO−<sub>2</sub>) and nitrate (NO<sup>−</sup><sub>3</sub>), the two stable products of NO. A two-way ANOVA with Bonferroni’s multiple comparison tests was used, <span class="html-italic">n</span> = 3, **; <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, n.s; not significant.</p>
Full article ">Figure 4
<p>PA_EXT administration diminishes the oxidative stress caused by RTN treatment in rat primary astrocytes. Rat primary astrocytes were treated with 3 μM RTN or 1 μg/mL PA_EXT. (<b>A</b>) Cellular ROS (CellROX) levels were measured using fluorescence spectrophotometry. CellROX intensity was normalized to Hoechst33342 intensity and compared to controls (DMSO alone). (<b>B</b>) NO<sup>−</sup>₂ and NO<sup>−</sup><sub>3</sub> levels were measured using the Griess assay and graphed. A two-way ANOVA with Bonferroni’s multiple comparison test was used, <span class="html-italic">n</span> = 3, *; <span class="html-italic">p</span> &lt; 0.05, ****; <span class="html-italic">p</span> &lt; 0.0001, n.s; not significant.</p>
Full article ">Figure 5
<p>Blockage of α-synuclein aggregation by PA_EXT. α-synuclein fibrilization was examined using 1 μg/mL recombinant monomeric α-synuclein with 50 ng/mL α-synuclein fibril seeds or without α-synuclein fibril (no fibril seed), with DMSO as a vehicle control or 1 μg/mL PA_EXT. (<b>A</b>) The thioflavin T assay was used to quantify the β-sheet structures, which are abundant in fibrillar α-synuclein. PBS was tested as a control in the Thioflavin T assay. (<b>B</b>) Western blotting was used to assess α-synuclein aggregation. (<b>C</b>) Densitometry was used to quantify the high molecular weight of α-synuclein (HWM) and monomeric α-synuclein (Mono). A one-way ANOVA and Bonferroni’s multiple comparison tests were used, n = 1, duplication assay, ***; <span class="html-italic">p</span> &lt; 0.001, ****; <span class="html-italic">p</span> &lt; 0.0001, ns; not significant. SH-SY5Y cells were differentiated for 7 d, and DMSO or PA_EXT were administered for 7 d. (<b>D</b>) The supernatants of cell lysate were subjected to sandwich ELISA for estimation of total or fibrillar α-synuclein levels. A two-way ANOVA and Sidak’s multiple comparison tests were used, <span class="html-italic">n</span> = 5. ***; <span class="html-italic">p</span> &lt; 0.001, ****; <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Elevation of lysosomal activity by PA_EXT in dSH cells. (<b>A</b>) Lysosomes in dSH cells were stained with LysoTracker Blue DND-22, and GCase activity was represented by the intensity of PFB-FDGlu, which is a substrate of glucocerebrosidase. All cells were also stained with SYTO 59 Red Fluorescent Nucleic Acid Stain (SYTO 59 Red). Cell images were captured at 460× optical zoom, and the right upper white bar on the PFB-FDGlu images (middle panels) indicates the scale bar (40 µm). (<b>B</b>) Lysosomal populations were measured via Lyso Tracker Blue DND-22 intensity, and GCase activity was estimated using PFB-FDGlu intensity. Fluorescence intensities were normalized to SYTO59 Red intensity. (<b>C</b>) The ratio of GCase activity to lysosomal population. (<b>D</b>) Cathepsin D activity was estimated in cell lysates. Student’s <span class="html-italic">t</span>-test was used, <span class="html-italic">n</span> = 4, ****; <span class="html-italic">p</span> &lt; 0.0001, ns.; not significant.</p>
Full article ">Figure 7
<p><span class="html-italic">C. elegans</span> lifespan analyses post-PA_EXT treatment. <span class="html-italic">C. elegans</span> co-supplemented with PA_EXT and RTN showed a marginal increase in mean lifespan compared to RTN-treatment alone. Life span analysis for each treatment was conducted once, with 100 worms used in each treatment group.</p>
Full article ">
28 pages, 8501 KiB  
Article
The Bifunctional Dimer Caffeine-Indan Attenuates α-Synuclein Misfolding, Neurodegeneration and Behavioral Deficits after Chronic Stimulation of Adenosine A1 Receptors
by Elisabet Jakova, Omozojie P. Aigbogun, Mohamed Taha Moutaoufik, Kevin J. H. Allen, Omer Munir, Devin Brown, Changiz Taghibiglou, Mohan Babu, Chris P. Phenix, Ed S. Krol and Francisco S. Cayabyab
Int. J. Mol. Sci. 2024, 25(17), 9386; https://doi.org/10.3390/ijms25179386 - 29 Aug 2024
Viewed by 757
Abstract
We previously found that chronic adenosine A1 receptor stimulation with N6-Cyclopentyladenosine increased α-synuclein misfolding and neurodegeneration in a novel α-synucleinopathy model, a hallmark of Parkinson’s disease. Here, we aimed to synthesize a dimer caffeine-indan linked by a 6-carbon chain to cross [...] Read more.
We previously found that chronic adenosine A1 receptor stimulation with N6-Cyclopentyladenosine increased α-synuclein misfolding and neurodegeneration in a novel α-synucleinopathy model, a hallmark of Parkinson’s disease. Here, we aimed to synthesize a dimer caffeine-indan linked by a 6-carbon chain to cross the blood–brain barrier and tested its ability to bind α-synuclein, reducing misfolding, behavioral abnormalities, and neurodegeneration in our rodent model. Behavioral tests and histological stains assessed neuroprotective effects of the dimer compound. A rapid synthesis of the 18F-labeled analogue enabled Positron Emission Tomography and Computed Tomography imaging for biodistribution measurement. Molecular docking analysis showed that the dimer binds to α-synuclein N- and C-termini and the non-amyloid-β-component (NAC) domain, similar to 1-aminoindan, and this binding promotes a neuroprotective α-synuclein “loop” conformation. The dimer also binds to the orthosteric binding site for adenosine within the adenosine A1 receptor. Immunohistochemistry and confocal imaging showed the dimer abolished α-synuclein upregulation and aggregation in the substantia nigra and hippocampus, and the dimer mitigated cognitive deficits, anxiety, despair, and motor abnormalities. The 18F-labeled dimer remained stable post-injection and distributed in various organs, notably in the brain, suggesting its potential as a Positron Emission Tomography tracer for α-synuclein and adenosine A1 receptor in Parkinson’s disease therapy. Full article
Show Figures

Figure 1

Figure 1
<p>Radiosynthesis of the <sup>18</sup>F-C<sub>8</sub>–6–I dimer and in vivo studies with the CD-1 mice. (<b>a</b>) Formation reaction of <sup>18</sup>F-C<sub>8</sub>–6–I from C<sub>8</sub>–6–I–OMs in 23 ± 5% rcy (decay corrected). (<b>b</b>) All steps involved in the (<b>b1</b>) radiosynthesis of <sup>18</sup>F-C<sub>8</sub>–6–I from [<sup>18</sup>F]fluorine purification, (<b>b2</b>) nucleophilic reaction with Kryptofix/K<sup>18</sup>F and (<b>b3</b>) semi-preparative HPLC purification, to (<b>b4</b>) PET-imaging and (<b>b5</b>) biodistribution of the <sup>18</sup>F- C<sub>8</sub>–6–I in major organs. (<b>b6</b>) The data were then analyzed and graphed using GraphPad Prism 8 (San Diego, CA, USA). Created using <a href="http://BioRender.com" target="_blank">BioRender.com</a> (URL accessed on 8 July 2024).</p>
Full article ">Figure 2
<p>Behavioral tests conducted for male Sprague-Dawley rats after the 7-day chronic injection of C<sub>8</sub>–6–I at 3 and 5 mg/kg. (<b>a</b>). Y-maze test values of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) as percentage of time spent in each of the arms: S-arm (“start” arm), O-arm (“old” arm), and N-arm (“new’ arm). The percentages of the time spent in each arm were calculated from the 5 min trial. Open field test values of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) as percentage of time spent in the center square. The animal is placed in the center square of the grid and left free to explore the field for 10 min. (<b>b</b>). The percentage of the time spent in the red center square. (<b>c</b>). The total fecal boli count. Forced swim test results of the 7-day chronic C<sub>8</sub>–6–I dimer (3 mg/kg and 5 mg/kg) treatment groups as measurements of swimming vigorously and successfully. The animals were placed in the forced swim tank and let free to swim for 10 min. Once the test was conducted the animals were scored for (<b>d</b>) vigor, the ability to purposely swim and use all limbs and (<b>e</b>) success, the ability to keep their head above water. (<b>f</b>). The total time spent immobile was also measured to assess learned helplessness and despair. Each dimer treatment was repeated at least 10 times per treatment (<span class="html-italic">n</span> = 10) and the average of each treatment is presented in bar graphs as means ± SEM. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison tests with * <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; 0.001.</p>
Full article ">Figure 3
<p>Molecular docking simulation of α-Syn structures ((<b>a</b>). C1, (<b>b</b>). C2, (<b>c</b>). C3, and (<b>d</b>). C8) bound to C<sub>8</sub>–6–I. Bold black dashed lines and amino acid residues indicate hydrogen bonding, while the grey dashed lines and amino acid residues indicate hydrophobic interactions. (<b>a</b>) The bifunctional dimer compound forms a hydrogen bond with T81 located in the NAC region of α-Syn, and additional hydrophobic interactions are found with the α-Syn N-terminus and NAC region. (<b>b</b>) The dimer compound interacts via hydrogen bonding to the α-Syn N-terminus, and additional hydrophobic interactions occur with amino acid residues in the N-terminus and NAC region. (<b>c</b>) The dimer compound binds via hydrogen bonds to amino acid residues in the α-Syn N-terminus and NAC region, and additional hydrophobic binding occurs with amino acids located in the distal N-terminus and NAC region. (<b>d</b>) The dimer compound binds via hydrogen bonding with NAC amino acid residues (V66, T72) and through hydrophobic interactions with amino acid residues in the N-terminus and NAC domain. C1, C2, C3 and C8 α-Syn structures bind to the dimer compound, which is predicted to form a “loop” conformation of α-Syn.</p>
Full article ">Figure 4
<p>Molecular docking of C<sub>8</sub>–6–I with A1R and A2AR. (<b>a</b>) Amino acid sequence alignment of human A1R and A2AR with distinct binding of C<sub>8</sub>–6–I to A1R and A2AR indicated (red asterisks, A1R binding; blue asterisks, A2AR binding). Amino acid residues shaded in red are conserved or identical amino acid sequences, while amino acids in red font are mostly classified under non-polar aliphatic residues (AVLIM). Other amino acids highlighted in red font are classified as follows: HKR are polar positive; DE are polar negative; STNQ are polar neutral; FYW are nonpolar aromatic. (<b>b</b>) Molecular docking of C<sub>8</sub>–6–I with A1R showing binding to amino acid residues that are similarly found within the A1R orthosteric binding site for adenosine. (<b>c</b>) Molecular docking showing C<sub>8</sub>–6–I binding to amino acid residues that do not resemble those associated with A2AR orthosteric binding site.</p>
Full article ">Figure 5
<p>Summary of the surface area analysis of the dimer study’s pars compacta region of the substantia nigra of DAPI, TH, and α-Syn. (<b>a</b>) Representative images of 40 μm pars compacta region of substantia nigra taken with 63X oil immersion objective of a confocal microscope (126 times magnification). Separate channels of 7-day chronic intraperitoneal injections with 3 mg/kg of the following treatments: Control (DMSO/Saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Slices were probed for DAPI (Blue), Thioflavin S (Thio-S, green), and α-Syn (Red, Alexa Fluor 647). Arrows indicate neuronal somas and processes with high localization of aggregated α-Syn. Scale 50 μm. (<b>b</b>) Bar charts showing the mean area intensities of α-Syn and Thioflavin S in the pars compacta region of the substantia nigra. Similar areas of 100-by-100 μm ROI coordinates for lateral pars compacta of SN were quantified, respectively, for each slice and normalized by subtracting F0 (50 by 50 μm ROI coordinates) values of the background (non-cell body bottom area). The average intensity values in bars represent the average mean ± SEM from <span class="html-italic">n</span> = 4 independent experiments. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison tests with ** <span class="html-italic">p</span> &lt; 0.01; and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>Summary of the surface area analysis of the CA1 region of the hippocampus of DAPI, α-Syn and Thio-S. (<b>a</b>) Representative images from 40 μm hippocampal rat brain slices after probing for DAPI, anti-α-Synuclein and Thioflavin S (Thio-S) taken at 63-times magnification with a confocal microscope for the following treatments: Control (DMSO/Saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Arrows indicate high colocalization of Thioflavin S and α-Syn in CA1 hippocampal neuronal somas and dendritic processes. Scale 50 μm. (<b>b</b>) Bar charts showing the mean area intensities of α-Syn and Thioflavin S in the CA1 region of the hippocampus. Fluorescence intensities from a 100-by-100 μm ROI from the CA1 pyramidal cell layer of the hippocampus were quantified using a similar method to that employed for the pars compacta region. The average intensity values in bars represent the average mean ± SEM from <span class="html-italic">n</span> = 4 independent experiments. Significances were determined using One-way ANOVA, followed by Student–Newman–Keuls multiple comparison test with * <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; 0.001.</p>
Full article ">Figure 7
<p>Fluoro-Jade C (FJC) staining in the SN pars compacta and in the hippocampus CA1 region of rats with 7-day chronic intraperitoneal injection of Control (DMSO/saline), CPA, C<sub>8</sub>–6–I (3 mg/kg) + CPA, and C<sub>8</sub>–6–I (5 mg/kg) + CPA. Representative images with 50 μm scale bar for the (<b>a</b>) SN pars compacta and (<b>b</b>) the CA1 region of the hippocampus. FJC fluorescence intensity in a 100 × 100 μm<sup>2</sup> region was normalized to the control group (100%). Values are shown as mean ± SEM. The average FJC fluorescence values were obtained from <span class="html-italic">n</span> = 4 independent experiments. * <span class="html-italic">p</span> &lt; 0.05; and ** <span class="html-italic">p</span> &lt; 0.01 (one-way ANOVA followed by Student–Newman–Keuls post-hoc multiple comparison test).</p>
Full article ">Figure 8
<p>Ex vivo stability and biodistribution in CD-1 mice at five different time points (5, 10, 20, 40, and 60 min) and major organs. (<b>a</b>) HPLC co-registration profiles of <sup>18</sup>F–C<sub>8</sub>–6–I and <sup>19</sup>F–C<sub>8</sub>–6–I based on a radio detector and ultraviolet detector, respectively. Analytical radio-HPLC chromatograms in mouse (<b>b</b>) liver and (<b>c</b>) pancreas extracts at 40 min after injection of <sup>18</sup>F–C<sub>8</sub>–6–I. (<b>d</b>) The distribution of <sup>18</sup>F–C<sub>8</sub>–6–I was calculated as a percentage of the injected dose per gram of tissue (% ID/g) and the results are displayed into two groups: lower distribution—blood, heart, bone, and brain (left panel), and higher distribution—liver, duodenum, kidneys, spleen, lungs, and large intestine (right panel). The data were obtained from <span class="html-italic">n</span> = 3 independent animals. Values are presented as mean ± SEM. Significances are indicated as follows: * <span class="html-italic">p</span> &lt; 0.05; and *** <span class="html-italic">p</span> &lt; 0.001 (one-way ANOVA followed by Student–Newman–Keuls post-hoc multiple comparison test).</p>
Full article ">Figure 9
<p>Representative PET/CT images in CD-1 mice at different time points throughout an hour, as well as time–activity curve in the brain. (<b>a</b>) PET images at different time points (1 min, 5 min, 10 min, 20 min, 30 min, and 50 min). (<b>b</b>) PET summation images for 60 min and (<b>c</b>) 120 min dynamic imaging. (<b>d</b>) Time–activity curve (TAC) of <sup>18</sup>F–C<sub>8</sub>–6–I for whole brain and three regions consisting of the cortex, midbrain, and hippocampus from PET/CT imaging. Values are presented as the standardized uptake value (SUV). The data were obtained from <span class="html-italic">n</span> = 3 independent animals.</p>
Full article ">
13 pages, 774 KiB  
Review
Active Immunotherapy for the Prevention of Alzheimer’s and Parkinson’s Disease
by Madeline M. Vroom and Jean-Cosme Dodart
Vaccines 2024, 12(9), 973; https://doi.org/10.3390/vaccines12090973 - 28 Aug 2024
Viewed by 1485
Abstract
Neurodegenerative diseases (ND) give rise to significant declines in motor, autonomic, behavioral, and cognitive functions. Of these conditions, Alzheimer’s disease (AD) and Parkinson’s disease (PD) are the most prevalent, impacting over 55 million people worldwide. Given the staggering financial toll on the global [...] Read more.
Neurodegenerative diseases (ND) give rise to significant declines in motor, autonomic, behavioral, and cognitive functions. Of these conditions, Alzheimer’s disease (AD) and Parkinson’s disease (PD) are the most prevalent, impacting over 55 million people worldwide. Given the staggering financial toll on the global economy and their widespread manifestation, NDs represent a critical issue for healthcare systems worldwide. Current treatment options merely seek to provide symptomatic relief or slow the rate of functional decline and remain financially inaccessible to many patients. Indeed, no therapy has yet demonstrated the potential to halt the trajectory of NDs, let alone reverse them. It is now recognized that brain accumulation of pathological variants of AD- or PD-associated proteins (i.e., amyloid-β, Tau, α-synuclein) begins years to decades before the onset of clinical symptoms. Accordingly, there is an urgent need to pursue therapies that prevent the neurodegenerative processes associated with pathological protein aggregation long before a clinical diagnosis can be made. These therapies must be safe, convenient, and affordable to ensure broad coverage in at-risk populations. Based on the need to intervene long before clinical symptoms appear, in this review, we present a rationale for greater investment to support the development of active immunotherapy for the prevention of the two most common NDs based on their safety profile, ability to specifically target pathological proteins, as well as the significantly lower costs associated with manufacturing and distribution, which stands to expand accessibility to millions of people globally. Full article
(This article belongs to the Special Issue Vaccine Coverage and Safety in Immunization Programs)
Show Figures

Figure 1

Figure 1
<p>Aggregate-prone variants of endogenous proteins are associated with ND. (<b>A</b>) In Alzheimer’s Disease, pathological Aβ and hyperphosphorylated Tau proteins form amyloid plaques and fibrillary tangles, respectively. (<b>B</b>) In Parkinson’s Disease, αSyn oligomerizes into protofibrils that aggregate further to form intracellular inclusions. Image created with BioRender.</p>
Full article ">
26 pages, 11651 KiB  
Article
The GBA1 K198E Variant Is Associated with Suppression of Glucocerebrosidase Activity, Autophagy Impairment, Oxidative Stress, Mitochondrial Damage, and Apoptosis in Skin Fibroblasts
by Laura Patricia Perez-Abshana, Miguel Mendivil-Perez, Marlene Jimenez-Del-Rio and Carlos Velez-Pardo
Int. J. Mol. Sci. 2024, 25(17), 9220; https://doi.org/10.3390/ijms25179220 - 25 Aug 2024
Viewed by 1014
Abstract
Parkinson’s disease (PD) is a multifactorial, chronic, and progressive neurodegenerative disorder inducing movement alterations as a result of the loss of dopaminergic (DAergic) neurons of the pars compacta in the substantia nigra and protein aggregates of alpha synuclein (α-Syn). Although its etiopathology agent [...] Read more.
Parkinson’s disease (PD) is a multifactorial, chronic, and progressive neurodegenerative disorder inducing movement alterations as a result of the loss of dopaminergic (DAergic) neurons of the pars compacta in the substantia nigra and protein aggregates of alpha synuclein (α-Syn). Although its etiopathology agent has not yet been clearly established, environmental and genetic factors have been suggested as the major contributors to the disease. Mutations in the glucosidase beta acid 1 (GBA1) gene, which encodes the lysosomal glucosylceramidase (GCase) enzyme, are one of the major genetic risks for PD. We found that the GBA1 K198E fibroblasts but not WT fibroblasts showed reduced catalytic activity of heterozygous mutant GCase by −70% but its expression levels increased by 3.68-fold; increased the acidification of autophagy vacuoles (e.g., autophagosomes, lysosomes, and autolysosomes) by +1600%; augmented the expression of autophagosome protein Beclin-1 (+133%) and LC3-II (+750%), and lysosomal–autophagosome fusion protein LAMP-2 (+107%); increased the accumulation of lysosomes (+400%); decreased the mitochondrial membrane potential (∆Ψm) by −19% but the expression of Parkin protein remained unperturbed; increased the oxidized DJ-1Cys106-SOH by +900%, as evidence of oxidative stress; increased phosphorylated LRRK2 at Ser935 (+1050%) along with phosphorylated α-synuclein (α-Syn) at pathological residue Ser129 (+1200%); increased the executer apoptotic protein caspase 3 (cleaved caspase 3) by +733%. Although exposure of WT fibroblasts to environmental neutoxin rotenone (ROT, 1 μM) exacerbated the autophagy–lysosomal system, oxidative stress, and apoptosis markers, ROT moderately increased those markers in GBA1 K198E fibroblasts. We concluded that the K198E mutation endogenously primes skin fibroblasts toward autophagy dysfunction, OS, and apoptosis. Our findings suggest that the GBA1 K198E fibroblasts are biochemically and molecularly equivalent to the response of WT GBA1 fibroblasts exposed to ROT. Full article
(This article belongs to the Special Issue Autophagy in Health, Aging and Disease, 4th Edition)
Show Figures

Figure 1

Figure 1
<p>Enzyme activity and expression levels of glucocerebrosidase (GCase) in fibroblasts bearing the mutation GBA1 K198E with in silico molecular docking of glucosylsphingosine (GlcSph) and GCase. (<b>A</b>) Enzyme activity of GCase activity in WT GBA1 (blue bar) and GBA1 K198E fibroblasts (red bar). (<b>B</b>) Protein expression levels of glucocerebrosidase (GCase) in WT GBA1 (blue curve) and GBA1 K198E fibroblasts (red curve) assessed by flow cytometry. (<b>C</b>) Quantification of GCase expression levels. Numbers in histograms represent positive cellular population for the tested marker. (<b>D</b>) Representative CB-Dock2 3D images showing the molecular docking of WT GCase (created by Alphafold2) with GlcSph (PubChem CID: 5280570). (<b>E</b>) Representative enlarged image of (<b>D</b>) showing the molecular docking of WT GCase with GlcSph. (<b>F</b>) Two-dimensional diagram showing conventional hydrogen bond between GCaseGlcSph interaction. (<b>G</b>) Representative CB-Dock2 3D images showing the molecular docking of WT GCase with conduritol-B-epoxide (CBE, CID: 136345). (<b>H</b>) Representative enlarged image of (G) showing the molecular docking of WT GCase with CBE. The data are presented as mean ± SD of two independent experiments in triplicated (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>GBA1 K198E variant induces acute GCase deficiency in the autophagy–lysosome system reflected as acidification of autophagosomes, lysosomes, and autolysosomes in untreated or treated fibroblast with rapamycin (RAP) or bafilomycin A1 (BAF). (<b>A</b>) Representative flow cytometry histograms showing the autophagy–lysosome acidification in untreated WT (blue curve) and GBA1 K198E fibroblasts (red curve), (<b>B</b>) WT and GBA1 K198E fibroblasts treated with rapamycin (RAP, 10 nM) or (<b>C</b>) bafilomycin A1 (BAF, 10 nM). (<b>D</b>) Quantitative analysis of autophagy–lysosome (acidification)-positive cells. (<b>E</b>–<b>G</b>) Representative immunofluorescence images showing autophagy–lysosome acidification in (<b>E</b>) untreated WT fibroblasts, (<b>F</b>) treated with rapamycin (RAP, 10 nM) or (<b>G</b>) treated bafilomycin A1 (BAF, 10 nM). (<b>H</b>–<b>J</b>) Representative immunofluorescence images showing autophagy–lysosome acidification in (<b>H</b>) untreated GBA1 K198E fibroblasts, (<b>I</b>) treated with rapamycin (RAP, 10 nM) or (<b>J</b>) treated bafilomycin A1 (BAF, 10 nM). (<b>K</b>) Quantitative analysis of autophagy lysosome as mean fluorescence intensity (MFI). Numbers in histograms represent positive cellular population for the tested marker. The histograms and photomicrographs represent 1 out of 3 independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; ns = not significant. Image magnification: 200×.</p>
Full article ">Figure 3
<p>GBA1 K198E variant upregulates expression of autophagic Beclin-1, LC3-II, and LAMP-2 proteins in fibroblasts. (<b>A</b>) Representative flow cytometry histogram analysis showing Beclin-1 expression in WT-GBA1 (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Quantitative (%) analysis of Beclin-1 expression in WT (<b>blue bar</b>) and GBA1 K198E fibroblast (red bar); (<b>C</b>) representative immunofluorescence image showing Beclin-1 reactivity in WT GBA1 and (<b>D</b>) K198E GBA1 fibroblasts (red fluorescence). (<b>E</b>) Quantitative (MFI) analysis of Beclin-1. (<b>F</b>) Representative flow cytometry histogram analysis showing LC3-II expression in WT GBA1 (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>G</b>) Quantitative analysis of LC3-II in WT (blue bar) and GBA1 K198E fibroblasts (red bar). (<b>H</b>) Representative immunofluorescence images showing LCIII-2 reactivity in fibroblasts WT-GBA1 and (<b>I</b>) GBA1 K198E fibroblasts (red fluorescence). (<b>J</b>) Quantitative analysis of LC3-II. (<b>K</b>) Representative flow cytometry histogram analysis showing LAMP-2 expression in WT-GBA1 (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>L</b>) Quantitative analysis of LAMP-2 in WT (blue bar) and GBA1 K198E fibroblasts (red bar). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>M</b>) Representative immunofluorescence images showing LAMP-2 reactivity in fibroblasts WT GBA1 and (<b>N</b>) GBA1 K198E fibroblasts (red fluorescence). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>O</b>) Quantitative analysis of LAMP-2. Numbers in histograms represent positive cellular population for the tested marker. Histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: * <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. Image magnification; 200×.</p>
Full article ">Figure 4
<p>GBA1 K198E variant increases the accumulation of lysosomes and decreases the mitochondrial membrane potential (∆Ψm) in mutant fibroblasts, while rotenone aggravates the damage. (<b>A</b>) Representative flow cytometry histogram showing WT (blue curve) or GBA1 K198E fibroblasts (red curve) stained with Lysotracker<sup>®</sup>. (<b>B</b>) Representative flow cytometry histogram showing WT (blue curve) or GBA1 K198E fibroblasts (red curve) exposed to rotenone (ROT, 1 μM) and stained with Lysotracker<sup>®</sup>. (<b>C</b>) Percentage of Lysotracker<sup>®</sup> stain-positive cells in untreated or treated WT (blue bar) and GBA1 K198E fibroblasts (red bar) with ROT. (<b>D</b>) Representative flow cytometry histogram showing WT (blue curve) or GBA1 K198E fibroblasts (red curve) stained with Mitotracker<sup>®</sup>. (<b>E</b>) Representative flow cytometry histogram showing WT (blue curve) or GBA1 K198E fibroblasts (red curve) exposed to rotenone (ROT, 1 μM) stained with Mitotracker<sup>®</sup>. (<b>F</b>) Percentage of Mitotracker<sup>®</sup> stain-positive cells in untreated or treated WT (blue bar) and GBA1 K198E fibroblasts (red bar) with ROT. (<b>G</b>) Representative fluorescence microscopy image showing untreated WT fibroblasts stained with Lysotracker<sup>®</sup> and Mitotracker<sup>®</sup> (red fluorescence), (<b>G’</b>) close-up of image G (white line square); (<b>H</b>) representative fluorescence microscopy image showing untreated GBA1 K198E fibroblasts stained with Lysotracker<sup>®</sup> (green fluorescence) and Mitotracker<sup>®</sup> (red fluorescence). (<b>H’</b>) Close-up of image H (white line square); (<b>I</b>) representative fluorescence microscopy photograph showing WT fibroblasts treated with ROT (1 μM) and stained with Lysotracker<sup>®</sup> (green fluorescence) and Mitotracker<sup>®</sup> (red fluorescence), (<b>I’</b>) close-up of image I (white line square); (<b>J</b>) representative fluorescence microscopy photograph showing GBA1 K198E fibroblasts treated with ROT (1 μM) and stained with Lysotracker<sup>®</sup> (green fluorescence) and Mitotracker<sup>®</sup>. (<b>J’</b>) Close-up of image J (white line square). (<b>K</b>) Quantification of the Lysotracker<sup>®</sup> mean fluorescence intensity (MFI) in untreated or treated WT (blue bar) and GBA1 K198E fibroblasts (red bar) with ROT. (<b>L</b>) Quantification of the MitoTracker<sup>®</sup> high mean fluorescence intensity (MFI) in untreated or treated WT (blue bar) and GBA1 K198E fibroblasts (red bar) with ROT. Nuclei were stained with Hoechst 33342 (blue fluorescence). Histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: * <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. Image magnification, 200×. ns = not significant.</p>
Full article ">Figure 5
<p>GBA1 K198E fibroblasts show unchanged expression levels of Parkin and mitochondrial colocalization of Parkin and TOM20 proteins, but Parkin shows a tendency to increase and mitochondrial colocalization upon rotenone exposure. (<b>A</b>) Representative flow cytometry histogram analysis showing the expression of Parkin protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Representative flow cytometry histogram analysis showing the expression of Parkin protein in WT (blue) and GBA1 K198E fibroblasts (red) upon rotenone (ROT, 1 μM) exposure. (<b>C</b>) Quantitative analysis of parkin expression. (<b>D</b>–<b>G</b>) Representative fluorescence merge density images of colocalization of Parkin and the translocase of the outer membrane of mitochondria 20 (TOM20) proteins in (<b>D</b>) WT fibroblasts (white fluorescence), (<b>E</b>) GBA1 K198E fibroblasts, (<b>F</b>) WT fibroblasts treated with rotenone (ROT, 1 μM), and (<b>G</b>) GBA1 K198E fibroblasts exposed to ROT (1 μM). (<b>D’</b>–<b>G’</b>) Representative fluorescence merge images in layer colocalization of Parkin (<b>D’’</b>–<b>G’</b>’, green fluorescence) with TOM20 proteins (<b>D’’’</b>–<b>G”’</b>, red fluorescence). Nuclei were stained with Hoechst 33342 (blue, <b>F’’’’</b>–<b>G’”’</b>). (<b>H</b>) Quantification of the Parkin/mitochondria mean fluorescence intensity (MFI). Flow cytometry histograms represent one out of three conducted experiments. The results are reported as the mean ± standard deviation of 3 independent experiments (dots in bar). Fluorescence microphotographs represent one out of three experiments (n=3). A one-way ANOVA, followed by Tukey’s test, was conducted for statistical analysis. Statistically significant variations are indicated by * <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; ns = no significance.</p>
Full article ">Figure 6
<p>GBA1 K198E fibroblasts show an endogenously high percentage of oxidized DJ-1-Cys106-SOH into DJ-1 Cys106SO<sub>3</sub>. (<b>A</b>) Representative flow cytometry histogram analysis showing oxidized DJ-1 (Cys106-SO<sub>3</sub>) protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Representative flow cytometry histogram analysis showing oxidized DJ-1 (Cys106-SO<sub>3</sub>) protein in WT (blue color) and GBA1 K198E fibroblasts (red color) upon rotenone (ROT, 1 μM) exposure. Flow cytometry histograms represent one out of three conducted experiments. The results are reported as the mean ± standard deviation of 3 independent experiments. (<b>C</b>) Quantitative analysis of oxidized DJ-1 (Cys106-SO<sub>3</sub>) protein. (<b>D</b>) Representative immunofluorescence image showing oxidized DJ-1 protein (Cys106-SO<sub>3,</sub> green fluorescence) in WT GBA1 fibroblasts. (<b>E</b>) Representative immunofluorescence image showing oxidized DJ-1 protein (Cys106-SO<sub>3,</sub> green fluorescence) in GBA1 K198E fibroblasts. (<b>F</b>) Representative fluorescence microscopy image showing oxidized DJ-1 protein (Cys106-SO<sub>3,</sub> green fluorescence) in WT GBA1 fibroblast treated with ROT (1 μM). (<b>G</b>) Representative fluorescence microscopy image showing oxidized DJ-1 (Cys106-SO<sub>3,</sub> green fluorescence) protein in GBA1 K198E fibroblast treated with ROT (1 μM). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>H</b>) Quantitative analysis of oxidized DJ-1 protein (Cys106-SO<sub>3</sub>). Numbers in histograms represent positive cellular population for the tested marker. The histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; ns = not significance. Image magnification: 200×.</p>
Full article ">Figure 7
<p>GBA1 K198E fibroblasts show an endogenously high percentage of phosphorylated LRRK2 at Ser935. (<b>A</b>) Representative flow cytometry histogram analysis showing phosphorylated LRRK2 at Ser935 protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Representative flow cytometry histogram analysis showing pSer935 LRRK2 protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve) upon rotenone (ROT, 1 μM) exposure. (<b>C</b>) Quantitative analysis of pSer935 LRRK2 protein. (<b>D</b>) Representative immunofluorescence image showing pSer935 LRRK2 protein (green fluorescence) in WT GBA1 fibroblasts. (<b>E</b>) Representative immunofluorescence image showing pSer935 LRRK2 protein (green fluorescence) in GBA1 K198E fibroblasts. (<b>F</b>) Representative fluorescence microscopy image showing pSer935 LRRK2 protein (green fluorescence) in WT GBA1 fibroblast treated with ROT (1 μM). (<b>G</b>) Representative fluorescence microscopy image showing pSer935 LRRK2 protein (green fluorescence) protein in GBA1 K198E fibroblast treated with ROT (1 μM). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>H</b>) Quantitative analysis of pSer935 LRRK2 protein. Numbers in histograms represent positive cellular population for the tested marker. The histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: * <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. Image magnification: 200×.</p>
Full article ">Figure 8
<p>GBA1 K198E fibroblasts show an endogenously high percentage of phosphorylated α-Syn at Ser129. (<b>A</b>) Representative flow cytometry histogram analysis showing phosphorylated α-Syn at Ser129 protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Representative flow cytometry histogram analysis showing pSer129 α-Syn protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve) upon rotenone (ROT, 1 μM) exposure. (<b>C</b>) Quantitative analysis of pSer129 α-Syn protein. (<b>D</b>) Representative immunofluorescence image showing pSer129 α-Syn protein (green fluorescence) in WT GBA1 fibroblasts. (<b>E</b>) Representative immunofluorescence image showing pSer129 α-Syn protein (green fluorescence) in GBA1 K198E fibroblasts. (<b>F</b>) Representative fluorescence microscopy image showing pSer129 α-Syn protein (green fluorescence) in WT GBA1 fibroblast treated with ROT (1 μM). (<b>G</b>) Representative fluorescence microscopy image showing pSer129 α-Syn protein (green fluorescence) in GBA1 K198E fibroblast treated with ROT (1 μM). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>H</b>) Quantitative analysis of pSer129 α-Syn protein. Numbers in histograms represent positive cellular population for the tested marker. The histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001. Image magnification: 200×.</p>
Full article ">Figure 9
<p>GBA1 K198E fibroblasts show endogenously high cleaved caspase 3 (CC3) compared to WT fibroblasts. (<b>A</b>) Representative flow cytometry histogram analysis showing cleaved caspase 3 (CC3) protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve). (<b>B</b>) Representative flow cytometry histogram analysis showing CC3 protein in WT (blue curve) and GBA1 K198E fibroblasts (red curve) upon rotenone (ROT, 1 μM) exposure. (<b>C</b>) Quantitative analysis of CC3 protein. (<b>D</b>) Representative immunofluorescence image showing CC3 protein (green fluorescence) in WT GBA1 fibroblasts. (<b>E</b>) Representative immunofluorescence image showing CC3 protein (green fluorescence) in GBA1 K198E fibroblasts. (<b>F</b>) Representative fluorescence microscopy image showing CC3 protein (green fluorescence) in WT GBA1 fibroblast treated with ROT (1 μM). (<b>G</b>) Representative fluorescence microscopy image showing CC3 protein (green fluorescence) protein in GBA1 K198E fibroblast treated with ROT (1 μM). Nuclei were stained with Hoechst 33342 (blue fluorescence). (<b>H</b>) Quantitative analysis of CC3 protein. Numbers in histograms represent positive cellular population for the tested marker. The histograms and photomicrographs represent one out of three independent experiments (n = 3). The data are presented as mean ± SD of three independent experiments (dots in bar). One-way ANOVA followed by Tukey’s test. Statistically significant differences: * <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. Image magnification: 200×.</p>
Full article ">Figure 10
<p>Schematic representation of the effect of K198E GCase on the autophagy–lysosomal pathway and apoptosis cell death in skin fibroblasts. (<b>A</b>) <span class="html-italic">Autophagy–lysosomal pathway and K198E GCase.</span> In WT GBA1 fibroblasts, the autophagy process begins with the ensemble of the ULK1 complex (unc-51-like kinase, ULK; autophagy-related protein 13, ATG13; RB1-inducible coiled-coil protein 1, FIP200; ATG101) (<b>1</b>), which then triggers nucleation of the phagophore (<b>2</b>) by phosphorylating components of the class III PI3K complex, involving Class IIIPI3K, vacuolar protein sorting 34 (VPS34) and Beclin 1 (Bc 1), among other proteins. These actions lead to the attachment of the microtubule-associated protein light chain 3 (LC3-II) to the phagophore (<b>3</b>), which further expand and form a sealed double-membrane, forming the autophagosome (<b>4</b>). This last vacuole, helped by lysosomal-associated membrane protein (LAMP-2, <b>5</b>), fused with the lysosome (<b>6</b>) to form the autolysosome (<b>7</b>), where unwanted cytosolic material (damaged mitochondria, protein aggregates, GlcCer (black dots)) is eliminated and recycled. On the other hand, enzymatic alteration of GCase mainly by genetic mutations (e.g., K198E) in at least one of the alleles of GBA1 almost leads to the undigested substrate GlcCer. As a result, lysosomes accumulate, thereby affecting the production line of autophagosomes, and autolysosomes. Indeed, heterozygous K198E GCase induces abnormal upregulation of protein Beclin 1, LC3-II and LAMP-2, and provokes an abnormally high accumulation of autophagosomes, lysosomes, and autolysosomes. Overall, K198E GCase provokes a highly deficient autophagy–lysosomal pathway in skin fibroblasts. (<b>B</b>) <span class="html-italic">Apoptosis pathway and K198E GCase.</span> In parallel, WT GCase interacts the mitochondrial respiratory component complex I (<b>8</b>, upper panel), thereby preserving energy metabolism (e.g., high ∆Ψm) and mitochondrial and nuclei integrity (<b>14</b>, upper panel). On the contrary, malfunction of mitochondrial Complex I due to improper interactions with K198E GCase (<b>8</b>, lower panel) allows electrons to leak, which are taken by molecular dioxygen (O2). Reduction of oxygen ends up in the formation of anion superoxide radicals (O2.-), which then dismutate into H2O2 (<b>9</b>). As a messenger molecule, H2O2 oxidized DJ-1Cys106-SOH (DJ-1red, <b>10</b>) into DJ-1Cys106-SO3 (-<span class="html-italic">sulfonic</span> group, DJ-1oxi, <b>11</b>) and induces the activation of the IKK complex (<b>10</b>), which phosphorylates LRRK2 at residue Ser935 (<b>11</b>). Phosphorylated LRRK2 kinase phosphorylates in turn the following three main targets: DLP-1 (dynamin-like protein), αSyn at residue Ser129, and PRDX3 (<b>12</b>). These three proteins might induce or contribute to mitochondria depolarization (e.g., low ∆Ψm), thereby inducing activation of caspase 3 (CASP3) into cleaved caspase 3 (CC3, <b>13</b>). Lastly, CC3 induces the fragmentation of nuclei (<b>14</b>). All these markers constitute typical signs of apoptosis (<b>8</b>–<b>14</b>).</p>
Full article ">
24 pages, 1262 KiB  
Review
The Role of α-Synuclein in Etiology of Neurodegenerative Diseases
by Daria Krawczuk, Magdalena Groblewska, Jan Mroczko, Izabela Winkel and Barbara Mroczko
Int. J. Mol. Sci. 2024, 25(17), 9197; https://doi.org/10.3390/ijms25179197 - 24 Aug 2024
Viewed by 737
Abstract
A presynaptic protein called α-synuclein plays a crucial role in synaptic function and neurotransmitter release. However, its misfolding and aggregation have been implicated in a variety of neurodegenerative diseases, particularly Parkinson’s disease, dementia with Lewy bodies, and multiple system atrophy. Emerging evidence suggests [...] Read more.
A presynaptic protein called α-synuclein plays a crucial role in synaptic function and neurotransmitter release. However, its misfolding and aggregation have been implicated in a variety of neurodegenerative diseases, particularly Parkinson’s disease, dementia with Lewy bodies, and multiple system atrophy. Emerging evidence suggests that α-synuclein interacts with various cellular pathways, including mitochondrial dysfunction, oxidative stress, and neuroinflammation, which contributes to neuronal cell death. Moreover, α-synuclein has been involved in the propagation of neurodegenerative processes through prion-like mechanisms, where misfolded proteins induce similar conformational changes in neighboring neurons. Understanding the multifaced roles of α-synuclein in neurodegeneration not only aids in acquiring more knowledge about the pathophysiology of these diseases but also highlights potential biomarkers and therapeutic targets for intervention in alpha-synucleinopathies. In this review, we provide a summary of the mechanisms by which α-synuclein contributes to neurodegenerative processes, focusing on its misfolding, oligomerization, and the formation of insoluble fibrils that form characteristic Lewy bodies. Furthermore, we compare the potential value of α-synuclein species in diagnosing and differentiating selected neurodegenerative diseases. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of α-synuclein.</p>
Full article ">Figure 2
<p>Biological functions of α-synuclein.</p>
Full article ">Figure 3
<p>Formation of α-Syn oligomers and fibrils.</p>
Full article ">
Back to TopTop