Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (98)

Search Parameters:
Keywords = beta secretase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 4640 KiB  
Article
In Vivo and Computational Studies on Sitagliptin’s Neuroprotective Role in Type 2 Diabetes Mellitus: Implications for Alzheimer’s Disease
by Vasudevan Mani and Minhajul Arfeen
Brain Sci. 2024, 14(12), 1191; https://doi.org/10.3390/brainsci14121191 - 26 Nov 2024
Viewed by 421
Abstract
Background/Objectives: Diabetes mellitus (DM), a widespread endocrine disorder characterized by chronic hyperglycemia, can cause nerve damage and increase the risk of neurodegenerative diseases such as Alzheimer’s disease (AD). Effective blood glucose management is essential, and sitagliptin (SITG), a dipeptidyl peptidase-4 (DPP-4) [...] Read more.
Background/Objectives: Diabetes mellitus (DM), a widespread endocrine disorder characterized by chronic hyperglycemia, can cause nerve damage and increase the risk of neurodegenerative diseases such as Alzheimer’s disease (AD). Effective blood glucose management is essential, and sitagliptin (SITG), a dipeptidyl peptidase-4 (DPP-4) inhibitor, may offer neuroprotective benefits in type 2 diabetes mellitus (T2DM). Methods: T2DM was induced in rats using nicotinamide (NICO) and streptozotocin (STZ), and biomarkers of AD and DM-linked enzymes, inflammation, oxidative stress, and apoptosis were evaluated in the brain. Computational studies supported the in vivo findings. Results: SITG significantly reduced the brain enzyme levels of acetylcholinesterase (AChE), beta-secretase-1 (BACE-1), DPP-4, and glycogen synthase kinase-3β (GSK-3β) in T2DM-induced rats. It also reduced inflammation by lowering cyclooxygenase-2 (COX-2), prostaglandin E2 (PGE2), tumor necrosis factor-α (TNF-α), and nuclear factor-κB (NF-κB). Additionally, SITG improved oxidative stress markers by reducing malondialdehyde (MDA) and enhancing glutathione (GSH). It increased anti-apoptotic B-cell lymphoma protein-2 (Bcl-2) while reducing pro-apoptotic markers such as Bcl-2-associated X (BAX) and Caspace-3. SITG also lowered blood glucose levels and improved plasma insulin levels. To explore potential molecular level mechanisms, docking was performed on AChE, COX-2, GSK-3β, BACE-1, and Caspace-3. The potential binding affinity of SITG for the above-mentioned target enzymes were 10.8, 8.0, 9.7, 7.7, and 7.9 kcal/mol, respectively, comparable to co-crystallized ligands. Further binding mode analysis of the lowest energy conformation revealed interactions with the critical residues. Conclusions: These findings highlight SITG’s neuroprotective molecular targets in T2DM-associated neurodegeneration and its potential as a therapeutic approach for AD, warranting further clinical investigations. Full article
Show Figures

Figure 1

Figure 1
<p>The timeline of the drug treatment and the experiment schedule.</p>
Full article ">Figure 2
<p>Effect of diabetes and sitagliptin on body weight in rats over a 30-day treatment period (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Day-1 in Control; ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in SITG10; <span>$</span> <span class="html-italic">p</span> &lt; 0.05 vs. Day 1 in T2DM + SITG30.</p>
Full article ">Figure 3
<p>Effect of sitagliptin on blood glucose levels in diabetes-induced rats (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in T2DM + SITG10; ### <span class="html-italic">p</span> &lt; 0.001 vs. Day 1 in T2DM + SITG30.</p>
Full article ">Figure 4
<p>Effect of sitagliptin on plasma insulin levels in diabetes-induced rats (<span class="html-italic">n</span> = 6). Data are presented as mean ± SEM. *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; ### <span class="html-italic">p</span> &lt; 0.001 vs. T2DM.</p>
Full article ">Figure 5
<p>Effect of sitagliptin on enzyme activity in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) <span class="html-italic">AChE</span>, (<b>B</b>) <span class="html-italic">BACE-1</span>, (<b>C</b>) <span class="html-italic">DPP-4</span>, and (<b>D</b>) <span class="html-italic">GSK-3β</span>. Data are presented as mean ± SEM. *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 vs. T2DM.</p>
Full article ">Figure 6
<p>Effect of sitagliptin on inflammatory markers in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) <span class="html-italic">COX-2</span>, (<b>B</b>) PGE2, (<b>C</b>) TNF-α, and (<b>D</b>) NF-κB. Data are presented as mean ± SEM. * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; # <span class="html-italic">p</span> &lt; 0.05 and ## <span class="html-italic">p</span> &lt; 0.01 vs. T2DM.</p>
Full article ">Figure 7
<p>Effect of sitagliptin on oxidative and antioxidant markers in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) MDA, (<b>B</b>) GSH, and (<b>C</b>) Catalase. Data are presented as mean ± SEM. ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. Control; # <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 vs. T2DM; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01 vs. T2DM + SITG10.</p>
Full article ">Figure 8
<p>Effect of sitagliptin on apoptotic proteins in the brains of diabetes-induced rats (<span class="html-italic">n</span> = 6): (<b>A</b>) Bcl-2, (<b>B</b>) BAX, and (<b>C</b>) Caspace-3. Data are presented as mean ± SEM. * <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 vs. Control; # <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 vs. T2DM.</p>
Full article ">Figure 9
<p>Binding mode of SITG in the active site of <span class="html-italic">AChE</span>, <span class="html-italic">COX-2</span>, <span class="html-italic">GSK-3β</span>, <span class="html-italic">BACE-1</span>, and Caspace-3.</p>
Full article ">
27 pages, 1765 KiB  
Review
Potential Roles of Hypoxia-Inducible Factor-1 in Alzheimer’s Disease: Beneficial or Detrimental?
by Tsu-Kung Lin, Chi-Ren Huang, Kai-Jung Lin, Yi-Heng Hsieh, Shang-Der Chen, Yi-Chun Lin, A-Ching Chao and Ding-I Yang
Antioxidants 2024, 13(11), 1378; https://doi.org/10.3390/antiox13111378 - 11 Nov 2024
Viewed by 722
Abstract
The major pathological characteristics of Alzheimer’s disease (AD) include senile plaques and neurofibrillary tangles (NFTs), which are mainly composed of aggregated amyloid-beta (Aβ) peptide and hyperphosphorylated tau protein, respectively. The excessive production of reactive oxygen species (ROS) and neuroinflammation are crucial contributing factors [...] Read more.
The major pathological characteristics of Alzheimer’s disease (AD) include senile plaques and neurofibrillary tangles (NFTs), which are mainly composed of aggregated amyloid-beta (Aβ) peptide and hyperphosphorylated tau protein, respectively. The excessive production of reactive oxygen species (ROS) and neuroinflammation are crucial contributing factors to the pathological mechanisms of AD. Hypoxia-inducible factor-1 (HIF-1) is a transcription factor critical for tissue adaption to low-oxygen tension. Growing evidence has suggested HIF-1 as a potential therapeutic target for AD; conversely, other experimental findings indicate that HIF-1 induction contributes to AD pathogenesis. These previous findings thus point to the complex, even contradictory, roles of HIF-1 in AD. In this review, we first introduce the general pathogenic mechanisms of AD as well as the potential pathophysiological roles of HIF-1 in cancer, immunity, and oxidative stress. Based on current experimental evidence in the literature, we then discuss the possible beneficial as well as detrimental mechanisms of HIF-1 in AD; these sections also include the summaries of multiple chemical reagents and proteins that have been shown to exert beneficial effects in AD via either the induction or inhibition of HIF-1. Full article
(This article belongs to the Special Issue Oxidative Stress and Alzheimer’s Disease)
Show Figures

Figure 1

Figure 1
<p>The major pathologies in AD brains include deposition of extracellular Aβ plaques and intraneuronal neurofibrillary tangles (NFTs) mainly composed of hyperphosphorylated tau proteins. Excessive Aβ aggregation can trigger diverse mechanisms including excitotoxicity, oxidative stress with heightened ROS levels, mitochondrial dysfunction with compromised ATP production, aberrant cell cycle reentry with subsequent apoptosis, and activation of neurotoxic glial cells like microglia to trigger neuroinflammation; these effects together lead to the damage or even demise of the neurons. Tau belongs to the microtubule-associated protein (MAP) family that is vital for microtubule assembly and stabilization in neuronal axons. Hyperphosphorylated tau proteins not only compromise microtubule structures to disturb axonal transport but also aberrantly aggregate into NFTs, which also contribute to neuroinflammation and neuronal apoptosis. Excessive neuronal death ultimately results in brain atrophy in AD patients.</p>
Full article ">Figure 2
<p>Activation of HIF-1 and its biological functions. HIF-1 is a heterodimeric transcription factor consisting of an oxygen-sensitive alpha subunit (HIF-1α) and a constitutively expressed beta subunit (HIF-1β). Under normoxia, HIF-1α undergoes hydroxylation of the proline residues catalyzed by the prolyl hydroxylase (PHD), which requires molecular oxygen (O<sub>2</sub>). The hydroxylated HIF-1α is then recognized by the von Hippel–Lindau (VHL) protein and E3 ubiquitin ligase for ubiquitination and subsequent proteasomal degradation. Under hypoxia, low-oxygen tension interferes with PHD hydroxylation and disrupts the interaction between HIF-1α and VHL, thereby stabilizing HIF-1α for its accumulation to form the heterodimeric HIF-1α/β. Translocation of the HIF-1α/β complex into the nucleus, along with coactivators p300/CBP, then drive the expression of target genes containing the hypoxia-response element (HRE) sequences in their promoters. HIF-1-dependent gene expression is crucial for numerous cellular responses to adapt the tissues to hypoxic environments, such as promoting angiogenesis and regulating vascular tone, enhancing antioxidation, regulating glucose transport and reprogramming energy metabolisms, affecting apoptosis, and regulating immune responses.</p>
Full article ">Figure 3
<p>Multiple AD-related mechanisms, including cerebral hypoperfusion, oxidative stress, and neuroinflammation, may trigger activation of HIF-1 to exert either positive or negative impacts on AD progression. The beneficial effects triggered by HIF-1 include affecting energy metabolisms, promoting neuroprotection/neurorestoration, enhancing neurogenesis, and counteracting oxidative stress, together allowing the tissues to adapt to the hypoxic environment. The detrimental effects include enhancing BACE1 expression with heightened β-secretase activity to promote Aβ production, impairing brain microvascular functions, and triggering neuronal cell cycle reentry followed by apoptosis. Notably, several unclear effects of HIF-1 in AD deserve detailed investigation. These include modulating brain circulation/angiogenesis, regulating tau hyperphosphorylation, affecting microglial functions and neuroinflammation, controlling the activities of α-secretase, γ-secretase, PS1/2 functions, and even Aβ degradation.</p>
Full article ">
26 pages, 3159 KiB  
Review
Haploinsufficiency and Alzheimer’s Disease: The Possible Pathogenic and Protective Genetic Factors
by Eva Bagyinszky and Seong Soo A. An
Int. J. Mol. Sci. 2024, 25(22), 11959; https://doi.org/10.3390/ijms252211959 - 7 Nov 2024
Viewed by 699
Abstract
Alzheimer’s disease (AD) is a complex neurodegenerative disorder influenced by various genetic factors. In addition to the well-established amyloid precursor protein (APP), Presenilin-1 (PSEN1), Presenilin-2 (PSEN2), and apolipoprotein E (APOE), several other genes such as [...] Read more.
Alzheimer’s disease (AD) is a complex neurodegenerative disorder influenced by various genetic factors. In addition to the well-established amyloid precursor protein (APP), Presenilin-1 (PSEN1), Presenilin-2 (PSEN2), and apolipoprotein E (APOE), several other genes such as Sortilin-related receptor 1 (SORL1), Phospholipid-transporting ATPase ABCA7 (ABCA7), Triggering Receptor Expressed on Myeloid Cells 2 (TREM2), Phosphatidylinositol-binding clathrin assembly protein (PICALM), and clusterin (CLU) were implicated. These genes contribute to neurodegeneration through both gain-of-function and loss-of-function mechanisms. While it was traditionally thought that heterozygosity in autosomal recessive mutations does not lead to disease, haploinsufficiency was linked to several conditions, including cancer, autism, and intellectual disabilities, indicating that a single functional gene copy may be insufficient for normal cellular functions. In AD, the haploinsufficiency of genes such as ABCA7 and SORL1 may play significant yet under-explored roles. Paradoxically, heterozygous knockouts of PSEN1 or PSEN2 can impair synaptic plasticity and alter the expression of genes involved in oxidative phosphorylation and cell adhesion. Animal studies examining haploinsufficient AD risk genes, such as vacuolar protein sorting-associated protein 35 (VPS35), sirtuin-3 (SIRT3), and PICALM, have shown that their knockout can exacerbate neurodegenerative processes by promoting amyloid production, accumulation, and inflammation. Conversely, haploinsufficiency in APOE, beta-secretase 1 (BACE1), and transmembrane protein 59 (TMEM59) was reported to confer neuroprotection by potentially slowing amyloid deposition and reducing microglial activation. Given its implications for other neurodegenerative diseases, the role of haploinsufficiency in AD requires further exploration. Modeling the mechanisms of gene knockout and monitoring their expression patterns is a promising approach to uncover AD-related pathways. However, challenges such as identifying susceptible genes, gene–environment interactions, phenotypic variability, and biomarker analysis must be addressed. Enhancing model systems through humanized animal or cell models, utilizing advanced research technologies, and integrating multi-omics data will be crucial for understanding disease pathways and developing new therapeutic strategies. Full article
(This article belongs to the Special Issue Genetic Mutations in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Gain-of-function vs. loss-of-function mechanisms of <span class="html-italic">PSEN</span> dysfunctions, leading to neurodegeneration.</p>
Full article ">Figure 2
<p>Potential effects of SORL1 haploinsufficiency, based on cell and animal models.</p>
Full article ">Figure 3
<p>Possible effects of ABCA7 haploinsufficiency, based on experiments on cell and animal models.</p>
Full article ">Figure 4
<p>Potential effects of <span class="html-italic">TREM2</span> haploinsufficiency, based on animal and cell models.</p>
Full article ">Figure 5
<p>Possible mechanisms associated with AD haploinsufficiency.</p>
Full article ">Figure 6
<p>Potential beneficial effects of haploinsufficiency in AD.</p>
Full article ">
12 pages, 856 KiB  
Article
Evaluation of Selected Plant Phenolics via Beta-Secretase-1 Inhibition, Molecular Docking, and Gene Expression Related to Alzheimer’s Disease
by Tugba Uçar Akyürek, Ilkay Erdogan Orhan, F. Sezer Şenol Deniz, Gokcen Eren, Busra Acar and Alaattin Sen
Pharmaceuticals 2024, 17(11), 1441; https://doi.org/10.3390/ph17111441 - 28 Oct 2024
Viewed by 748
Abstract
Background: The goal of the current study was to investigate the inhibitory activity of six phenolic compounds, i.e., rosmarinic acid, gallic acid, oleuropein, epigallocatechin gallate (EGCG), 3-hydroxytyrosol, and quercetin, against β-site amyloid precursor protein cleaving enzyme-1 (BACE1), also known as β-secretase or memapsin [...] Read more.
Background: The goal of the current study was to investigate the inhibitory activity of six phenolic compounds, i.e., rosmarinic acid, gallic acid, oleuropein, epigallocatechin gallate (EGCG), 3-hydroxytyrosol, and quercetin, against β-site amyloid precursor protein cleaving enzyme-1 (BACE1), also known as β-secretase or memapsin 2, which is implicated in the pathogenesis of Alzheimer’s disease (AD). Methods and Results: The inhibitory potential against BACE1, molecular docking simulations, as well as neurotoxicity and the effect on the AD-related gene expression of the selected phenolics were tested. BACE1 inhibitory activity was carried out using the ELISA microplate assay via fluorescence resonance energy transfer (FRET) technology. Molecular docking experiments were performed in the human BACE1 active site (PDB code: 2WJO). Neurotoxicity of the compounds was carried out in SH-SY5Y, a human neuroblastoma cell line, by the Alamar Blue method. A gene expression analysis of the compounds on fourteen genes linked to AD was conducted using the real-time polymerase chain reaction (RT-PCR) method. Rosmarinic acid, EGCG, oleuropein, and quercetin (also used as the reference) were able to inhibit BACE1 with their respective IC50 values 4.06 ± 0.68, 1.62 ± 0.12, 9.87 ± 1.01, and 3.16 ± 0.30 mM. The inhibitory compounds were observed to occupy the non-catalytic site of the BACE1. However, hydrogen bonds were found to be present between rosmarinic acid and EGCG and aspartic amino acid D228 in the catalytic site. Oleuropein and quercetin effectively suppressed the expression of PSEN, APOE, and CLU, which are recognized to be linked to the pathogenesis of AD. Conclusions: The outcomes of the work bring quercetin, EGCG, and rosmarinic acid to the forefront as promising BACE1 inhibitors. Full article
Show Figures

Figure 1

Figure 1
<p>The predicted binding modes for rosmarinic acid (<b>A</b>), EGCG (<b>B</b>), oleuropein (<b>C</b>), and quercetin (<b>D</b>) in human BACE1 active site (PDB: 2WJO). The yellow dotted lines represent H-bonds, and the cyan-dotted line represents π-π interactions.</p>
Full article ">Figure 2
<p>Fold regulation values of rosmarinic acid, EGCG, oleuropein, quercetin, 3–hydroxytyrosol, and gallic acid on genes compared to the control group. Multiple groups were compared by ANOVA and Dunnet <span class="html-italic">post hoc</span> test. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
17 pages, 10469 KiB  
Article
Multi-Protective Effects of Petunidin-3-O-(trans-p-coumaroylrutinoside)-5-O-glucoside on D-Gal-Induced Aging Mice
by Ruinan Wang, Lichengcheng Ren, Yue Wang, Na Hu, Fangfang Tie, Qi Dong and Honglun Wang
Int. J. Mol. Sci. 2024, 25(20), 11014; https://doi.org/10.3390/ijms252011014 - 13 Oct 2024
Viewed by 1030
Abstract
Petunidin-3-O-(trans-p-coumaroylrutinoside)-5-O-glucoside (PtCG), the primary anthocyanin ingredient in Lycium ruthenicum Murr., possesses a range of biological activities, including antioxidative properties and melanin inhibition. This study aimed to investigate the protective effect of PtCG on D-galactose (D-gal)-induced aging in female mice and elucidate [...] Read more.
Petunidin-3-O-(trans-p-coumaroylrutinoside)-5-O-glucoside (PtCG), the primary anthocyanin ingredient in Lycium ruthenicum Murr., possesses a range of biological activities, including antioxidative properties and melanin inhibition. This study aimed to investigate the protective effect of PtCG on D-galactose (D-gal)-induced aging in female mice and elucidate the underlying molecular pathways. Behavioral experiments, including the MWW and Y-maze tests, revealed that PtCG significantly ameliorated cognitive decline and enhanced learning and memory abilities in aging mice. Regarding biochemical indicators, PtCG considerably improved superoxide dismutase (SOD) and glutathione (GSH) activity while reducing malondialdehyde (MDA) and acetylcholinesterase (AChE) levels in the hippocampus and serum. Furthermore, PtCG ingestion alleviated liver injury by decreasing alanine transaminase (ALT), aspartate transaminase (AST), and alkaline phosphatase (AKP) levels, and attenuated renal damage by reducing blood urea nitrogen (BUN) and uric acid (UA) levels. Transmission electron microscopy (TEM) results demonstrated that PtCG restored the function and quantity of synapses in the hippocampus. Hematoxylin and eosin (H&E), Masson’s trichrome, and Nissl staining revealed that PtCG significantly improved the relevant pathological characteristics of liver and hippocampal tissues in aging mice. The molecular mechanism investigation showed that PtCG downregulated the protein expression of microglial marker ionized calcium-binding adapter molecule 1 (Iba1), astrocytic marker glial fibrillary acidic protein (GFAP), β-secretase 1 (BACE-1), and amyloid-beta1–42 (Aβ1–42) in the hippocampus of aging mice. The protein expression of inflammatory pathway components, including nuclear factor-kappa B (NF-κB), cyclooxygenase-2 (COX2), inducible nitric oxide synthase (iNOS), and interleukin-1 beta (IL-1β), was also suppressed. These findings suggest that PtCG may possess anti-aging properties, with its mechanism of action potentially linked to the attenuation of neuroinflammation, oxidative stress, and liver and kidney damage. PtCG may have future applications as a functional food for the treatment of aging-related disorders. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Figure 1

Figure 1
<p>PtCG ameliorated cognitive dysfunction during the behavioral testing of aging mice: (<b>a</b>) total swimming paths of the respective groups on the sixth day. Heat map analysis of animal tracking following MWM test. Different colors indicate location preference. Red represents increased time spent and blue represents minimal time spent during trial. I–IV represent quadrant 1–quadrant 4, respectively. The quadrant in which the target platform is located is the target quadrant (quadrant 4). (<b>b</b>) The escape latency for five consecutive daily tests. (<b>c</b>) Swimming speed in the probe trial. (<b>d</b>) Time spent in the target quadrant of the probe trial. (<b>e</b>) The number of times mice swam across the target platform in the probe trial. (<b>f</b>) The total number of arm entries. (<b>g</b>) Percentage alternation in the Y-maze test. Data are expressed as mean ± SD (n = 8 per group). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. Con group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. D-gal group.</p>
Full article ">Figure 2
<p>PtCG regulated the MDA content and SOD, GSH, AChE activity in the serum and hippocampus of aging mice. (<b>a</b>) Activity of SOD in the serum. (<b>b</b>) Activity of GSH in the serum. (<b>c</b>) Content of MDA in the serum. (<b>d</b>) Activity of AChE in the serum. (<b>e</b>) Activity of SOD in the hippocampus. (<b>f</b>) Activity of GSH in the hippocampus. (<b>g</b>) Content of MDA in the hippocampus. (<b>h</b>) Activity of AChE in the hippocampus. Data are expressed as mean ± SD (n = 8 per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. D-gal group.</p>
Full article ">Figure 3
<p>PtCG attenuated AKP, ALT, and AST levels in the liver and BUN and UA levels in the kidney of aging mice. (<b>a</b>) Levels of AKP in the liver; (<b>b</b>) levels of AST in the liver; (<b>c</b>) levels of ALT in the liver; (<b>d</b>) levels of BUN in the kidney; (<b>e</b>) levels of UA in the kidney. Data are expressed as mean ± SD (n = 8 per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. D-gal group.</p>
Full article ">Figure 4
<p>PtCG attenuated neuronal damage in the hippocampus of aging mice. (<b>a</b>) Microstructure observation of hippocampus with TEM. Scale bar, 2 μm and 500 nm. The red letter represent each synapse. The green letters and numbers represent each synaptic vesicle and the thickness of PSD. (<b>b</b>) The number of synapses; (<b>c</b>) the number of synaptic vesicles; (<b>d</b>) the thickness of postsynaptic density; (<b>e</b>) the hippocampal of mice were observed using H&amp;E staining: magnification 40× and 80×, Neurons shrink (black arrow), neurons become denatured (yellow arrow); (<b>f</b>) the hippocampal of mice were observed using Nissl staining: magnification 40× and 80×. The cells atrophied with vacuoles (black arrows) and intercellular spaces (red Arrows). Data are expressed as mean ± SD (n = 3 per group).</p>
Full article ">Figure 4 Cont.
<p>PtCG attenuated neuronal damage in the hippocampus of aging mice. (<b>a</b>) Microstructure observation of hippocampus with TEM. Scale bar, 2 μm and 500 nm. The red letter represent each synapse. The green letters and numbers represent each synaptic vesicle and the thickness of PSD. (<b>b</b>) The number of synapses; (<b>c</b>) the number of synaptic vesicles; (<b>d</b>) the thickness of postsynaptic density; (<b>e</b>) the hippocampal of mice were observed using H&amp;E staining: magnification 40× and 80×, Neurons shrink (black arrow), neurons become denatured (yellow arrow); (<b>f</b>) the hippocampal of mice were observed using Nissl staining: magnification 40× and 80×. The cells atrophied with vacuoles (black arrows) and intercellular spaces (red Arrows). Data are expressed as mean ± SD (n = 3 per group).</p>
Full article ">Figure 5
<p>PtCG attenuated pathological damage in the liver of aging mice. (<b>a</b>) H&amp;E staining of liver: magnification 40× and 80×. (<b>b</b>) Masson staining of liver: magnification 20× and 40×. (<b>c</b>) Statistical analysis of the positive areas by Masson staining. Data are expressed as mean ± SD (n = 3 per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Con group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. D-gal group.</p>
Full article ">Figure 6
<p>PtCG downregulated the expression of related proteins in the hippocampus of aging mice. (<b>a</b>) Protein expression levels of Iba1 and GFAP; (<b>b</b>) protein expression levels of BACE-1 and Aβ<sub>1–42</sub>; (<b>c</b>) protein expression levels of the NF-κB inflammatory signaling pathway. Data are expressed as mean ± SD (n = 3 per group). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. Con group; <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01 vs. D-gal group.</p>
Full article ">Figure 7
<p>Schematic diagram of anti-aging mechanism of PtCG. Note: Red up arrows represent upregulation of expression. Grey dashed arrows represent the inhibitory effect. Solid arrows represent promoting effect. (The mechanism diagram was drawn by using Figdraw 2.0 (24 September 2024).)</p>
Full article ">Figure 8
<p>Experimental procedure and treatment schedule.</p>
Full article ">
26 pages, 5799 KiB  
Review
Exploring the Benzazoles Derivatives as Pharmacophores for AChE, BACE1, and as Anti-Aβ Aggregation to Find Multitarget Compounds against Alzheimer’s Disease
by Martha Cecilia Rosales Hernández, Marycruz Olvera-Valdez, Jazziel Velazquez Toledano, Jessica Elena Mendieta Wejebe, Leticia Guadalupe Fragoso Morales and Alejandro Cruz
Molecules 2024, 29(19), 4780; https://doi.org/10.3390/molecules29194780 - 9 Oct 2024
Viewed by 1181
Abstract
Despite the great effort that has gone into developing new molecules as multitarget compounds to treat Alzheimer’s disease (AD), none of these have been approved to treat this disease. Therefore, it will be interesting to determine whether benzazoles such as benzimidazole, benzoxazole, and [...] Read more.
Despite the great effort that has gone into developing new molecules as multitarget compounds to treat Alzheimer’s disease (AD), none of these have been approved to treat this disease. Therefore, it will be interesting to determine whether benzazoles such as benzimidazole, benzoxazole, and benzothiazole, employed as pharmacophores, could act as multitarget drugs. AD is a multifactorial disease in which several pharmacological targets have been identified—some are involved with amyloid beta (Aβ) production, such as beta secretase (BACE1) and beta amyloid aggregation, while others are involved with the cholinergic system as acetylcholinesterase (AChE) and butirylcholinesterase (BChE) and nicotinic and muscarinic receptors, as well as the hyperphosphorylation of microtubule-associated protein (tau). In this review, we describe the in silico and in vitro evaluation of benzazoles on three important targets in AD: AChE, BACE1, and Aβ. Benzothiazoles and benzimidazoles could be the best benzazoles to act as multitarget drugs for AD because they have been widely evaluated as AChE inhibitors, forming π–π interactions with W286, W86, Y72, and F338, as well as in the AChE gorge and catalytic site. In addition, the sulfur atom from benzothiazol interacts with S286 and the aromatic ring from W84, with these compounds having an IC50 value in the μM range. Also, benzimidazoles and benzothiazoles can inhibit Aβ aggregation. However, even though benzazoles have not been widely evaluated on BACE1, benzimidazoles evaluated in vitro showed an IC50 value in the nM range. Therefore, important chemical modifications could be considered to improve multitarget benzazoles’ activity, such as substitutions in the aromatic ring with electron withdrawal at position five, or a linker 3 or 4 carbons in length, which would allow for better interaction with targets. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structures of benzazoles. Benzothiazole (<b>a</b>), riluzole (<b>b</b>), thiourea (<b>c</b>), urea (<b>d</b>), dexpramipexole (<b>e</b>), and pramipexole (<b>f</b>). Derivative compounds from riluzole with tioguanidines compound <b>3b</b> (<b>g</b>) and compound <b>3d</b> (<b>h</b>); benzimidazole (<b>i</b>) and benzoxazole (<b>j</b>).</p>
Full article ">Figure 2
<p>Chemical structures of benzoxazole derivatives used as AChE and BChE inhibitors. Compounds <b>11</b> (<b>a</b>) [<a href="#B46-molecules-29-04780" class="html-bibr">46</a>], <b>3g</b> (<b>b</b>) [<a href="#B47-molecules-29-04780" class="html-bibr">47</a>], <b>34</b> (<b>c</b>) [<a href="#B48-molecules-29-04780" class="html-bibr">48</a>], <b>5f</b> (<b>d</b>) [<a href="#B49-molecules-29-04780" class="html-bibr">49</a>], <b>32a</b> (<b>e</b>) [<a href="#B50-molecules-29-04780" class="html-bibr">50</a>], <b>33</b> (<b>f</b>) [<a href="#B50-molecules-29-04780" class="html-bibr">50</a>], <b>1g</b> (<b>g</b>) [<a href="#B51-molecules-29-04780" class="html-bibr">51</a>], <b>1a</b> (<b>h</b>) [<a href="#B51-molecules-29-04780" class="html-bibr">51</a>], and <b>1d</b> (<b>i</b>) [<a href="#B51-molecules-29-04780" class="html-bibr">51</a>].</p>
Full article ">Figure 3
<p>Chemical structures of benzothiazole compounds used as AChE inhibitors. Compounds <b>10w</b> (<b>a</b>) [<a href="#B53-molecules-29-04780" class="html-bibr">53</a>], <b>7a</b> (<b>b</b>), <b>7b</b> (<b>c</b>), <b>7c</b> (<b>d</b>), <b>7d</b> (<b>e</b>), <b>7e</b> (<b>f</b>) [<a href="#B54-molecules-29-04780" class="html-bibr">54</a>], <b>A5</b> (<b>g</b>), <b>A13</b> (<b>h</b>) [<a href="#B55-molecules-29-04780" class="html-bibr">55</a>], and <b>BPCT</b> (<b>i</b>) [<a href="#B56-molecules-29-04780" class="html-bibr">56</a>].</p>
Full article ">Figure 4
<p>Chemical structures of compounds <b>3d</b> (<b>a</b>), <b>3h</b> (<b>b</b>) [<a href="#B61-molecules-29-04780" class="html-bibr">61</a>], <b>2e</b> (<b>c</b>), <b>3c</b> (<b>d</b>), <b>3e</b> (<b>e</b>) [<a href="#B62-molecules-29-04780" class="html-bibr">62</a>], <b>5IIc</b> (<b>f</b>) [<a href="#B63-molecules-29-04780" class="html-bibr">63</a>], <b>4b</b> (<b>g</b>) [<a href="#B64-molecules-29-04780" class="html-bibr">64</a>], <b>7e</b> (<b>h</b>), and of donepezil (<b>i</b>) [<a href="#B65-molecules-29-04780" class="html-bibr">65</a>].</p>
Full article ">Figure 5
<p>Chemical structures of compounds <b>A12</b> (<b>a</b>) [<a href="#B67-molecules-29-04780" class="html-bibr">67</a>], <b>A1</b> (<b>b</b>), <b>A2</b> (<b>c</b>), <b>A3</b> (<b>d</b>), <b>A4</b> (<b>e</b>) [<a href="#B68-molecules-29-04780" class="html-bibr">68</a>], <b>16</b> (<b>f</b>), <b>21</b> (<b>g</b>) [<a href="#B69-molecules-29-04780" class="html-bibr">69</a>], <b>11</b> (<b>h</b>) [<a href="#B70-molecules-29-04780" class="html-bibr">70</a>], <b>12d</b> (<b>i</b>), and <b>12k</b> (<b>j</b>) [<a href="#B71-molecules-29-04780" class="html-bibr">71</a>].</p>
Full article ">Figure 6
<p>Chemical structures of compounds <b>1b</b> (<b>a</b>), <b>1c</b> (<b>b</b>), <b>1g</b> (<b>c</b>), <b>2c</b> (<b>d</b>), <b>2e</b> (<b>e</b>), <b>2h</b> (<b>f</b>) [<a href="#B72-molecules-29-04780" class="html-bibr">72</a>], <b>15g</b> (<b>g</b>), <b>15b</b> (<b>h</b>) [<a href="#B73-molecules-29-04780" class="html-bibr">73</a>], <b>12</b> (<b>i</b>), and <b>13</b> (<b>j</b>) [<a href="#B74-molecules-29-04780" class="html-bibr">74</a>].</p>
Full article ">Figure 7
<p>Chemical structures of compounds <b>5</b> (<b>a</b>) [<a href="#B75-molecules-29-04780" class="html-bibr">75</a>], <b>11</b> (<b>b</b>), <b>14</b> (<b>c</b>) [<a href="#B76-molecules-29-04780" class="html-bibr">76</a>], <b>34</b> (<b>d</b>) [<a href="#B77-molecules-29-04780" class="html-bibr">77</a>], and <b>7c</b> (<b>e</b>) [<a href="#B78-molecules-29-04780" class="html-bibr">78</a>].</p>
Full article ">Figure 8
<p>Principal targets to be inhibited by benzazoles as possible multitarget drugs for the treatment of AD. Each row indicated the enzyme or peptide that it inhibited for each benzazole. R<sup>1</sup> indicated the substitution in the benzazole ring in the 2 position. Figure created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 9
<p>Compound <b>4f</b> (<b>a</b>) [<a href="#B93-molecules-29-04780" class="html-bibr">93</a>], compound <b>3d</b> (<b>b</b>) [<a href="#B95-molecules-29-04780" class="html-bibr">95</a>], compound <b>TAC-BIM1</b> (<b>c</b>), compound <b>TAC-BIM2</b> (<b>d</b>) [<a href="#B96-molecules-29-04780" class="html-bibr">96</a>], compound <b>4c</b> (<b>e</b>) and compound <b>4g</b> (<b>f</b>) [<a href="#B97-molecules-29-04780" class="html-bibr">97</a>].</p>
Full article ">
13 pages, 2383 KiB  
Article
Glutamate Transporter 1 as a Novel Negative Regulator of Amyloid β
by Priyanka Sinha, Yuliia Turchyna, Shane Patrick Clancy Mitchell, Michael Sadek, Gokce Armagan, Florian Perrin, Masato Maesako and Oksana Berezovska
Cells 2024, 13(19), 1600; https://doi.org/10.3390/cells13191600 - 24 Sep 2024
Viewed by 919
Abstract
Glutamate transporter-1 (GLT-1) dynamics are implicated in excitotoxicity and Alzheimer’s disease (AD) progression. Early stages of AD are often marked by hyperactivity and increased epileptiform activity preceding cognitive decline. Previously, we identified a direct interaction between GLT-1 and Presenilin 1 (PS1) in the [...] Read more.
Glutamate transporter-1 (GLT-1) dynamics are implicated in excitotoxicity and Alzheimer’s disease (AD) progression. Early stages of AD are often marked by hyperactivity and increased epileptiform activity preceding cognitive decline. Previously, we identified a direct interaction between GLT-1 and Presenilin 1 (PS1) in the brain, highlighting GLT-1 as a promising target in AD research. This study reports the significance of this interaction and uncovers a novel role of GLT-1 in modulating amyloid-beta (Aβ) production. Overexpression of GLT-1 in cells reduces the levels of Aβ40 and Aβ42 by decreasing γ-secretase activity pertinent to APP processing and induces a more “open” PS1 conformation, resulting in decreased Aβ42/40 ratio. Inhibition of the GLT-1/PS1 interaction using cell-permeable peptides produced an opposing effect on Aβ, highlighting the pivotal role of this interaction in regulating Aβ levels. These findings emphasize the potential of targeting the GLT-1/PS1 interaction as a novel therapeutic strategy for AD. Full article
(This article belongs to the Special Issue Research on the Amyloid in Alzheimer’s Disease)
Show Figures

Figure 1

Figure 1
<p>Overexpression reduces secreted Aβ in cells. Empty vector pcDNA or GLT-1 was overexpressed in CHO PS70 cells, and Aβ species were measured from the conditioned medium. Absolute concentrations of (<b>A</b>) Aβ40 and (<b>B</b>) Aβ42 are plotted in pmol/L, along with (<b>C</b>) the Aβ42/40 ratio; n = 6. Significant reduction in the above-mentioned Aβ species was observed along with a significant reduction in Aβ42/40 ratio after GLT-1 transfection as compared to those transfected with pcDNA. (<b>D</b>) Western blotting showing levels of APP full-length and CTFs in cells transfected with either pcDNA or GLT-1; n = 3. Quantification of band intensities shows (<b>E</b>) no significant change in APP full-length levels post GLT-1 transfection and (<b>F</b>) significant accumulation of APP CTFs in GLT-1-transfected cells as compared to cells transfected with pcDNA when a ratio of full-length APP CTFs/APP was performed. Statistical significance was calculated using unpaired t-test with Mann–Whitney test to compare ranks (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>GLT-1 overexpression leads to “open” PS1 conformation. PS1 conformation was analyzed after pcDNA/GLT-1 transfection in CHO PS70 cells using FLIM. (<b>A</b>) The cells were stained with PS1 N- and C-termini antibodies followed by Alexa fluor 488 and Cy3 fluorescent antibodies, respectively. The last panel shows pseudo-colored lifetime images representing the donor fluorophore lifetime in picoseconds. The blue–green pixels represent greater distance between the fluorescently labeled PS1 N- and PS1 C-termini, indicating an “open” conformation. (<b>B</b>) The analysis of FRET efficiency was used to estimate the relative change in proximity between PS1 N- and PS1 C-termini. The graph shows the percentage of FRET efficiency, depicted as a violin plot with median (solid bars) and 25th and 75th percentiles (dotted bars); n = 6 independent experiments. Statistical significance was assessed using unpaired <span class="html-italic">t</span>-test with Mann–Whitney test to compare ranks (** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 3
<p>GLToverexpression reduces APP C99 processing by γ-secretase. (<b>A</b>) The schematic illustrating the structure of the C99 YT biosensor. Endogenous PS/γ-secretase cleaves the C99 portion of the biosensor, leading to a reduction in FRET between YPet (Y; donor) and Turquoise-GL (T; acceptor), and thus revealing γ-secretase activity. (<b>B</b>) The graph shows a normalized 531/489 nm ratio reflecting FRET efficiency between the Y and T fluorescent moieties, depicted as a violin plot. The higher the 531/489 nm ratio, the lower the γ-secretase activity. The median value is shown by solid bars; n = 9 independent experiments. pcDNA-transfected cells treated with DAPT (γ-secretase inhibitor) served as a positive control. Statistical significance was determined using Kruskal–Wallis ANOVA with Dunn’s multiple comparison test (**** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 4
<p>Inhibition of the GLT-1 interaction with PS1 increases secreted Aβ in primary neurons. Primary neurons (12–14 DIV) were treated with 5 µM cell-permeable peptide (CPP) for 2 h, and Aβ species in the conditioned medium were measured by ELISA. After enriching the media for secreted Aβ, absolute concentrations of Aβ40 and Aβ42 were plotted in pmol/L, along with the Aβ42/40 ratio; n = 8. (<b>A</b>) Aβ40 levels after GLT-1 CPP and (<b>D</b>) PS1 CPP treatments. Absolute concentrations of Aβ42 are shown after (<b>B</b>) GLT-1 CPP and (<b>E</b>) PS1 CPP treatment. (<b>C</b>,<b>F</b>) Aβ42/40 ratios after GLT-1 CPP and PS1 CPP treatments, respectively. (<b>G</b>,<b>J</b>) Western blotting showing levels of APP full-length and CTFs in neurons treated with either GLT-1 or PS1 CPPs or their scrambled counterparts; n = 3. (<b>H</b>,<b>K</b>) Quantification of band intensities shows no significant change in APP full-length levels post-CPP treatment. (<b>I</b>,<b>L</b>) There was significant accumulation of APP CTFs in CPP-treated neurons as compared to neurons treated with their scrambled counterparts when a ratio of full-length APP CTFs/APP was calculated. Statistical significance was calculated using unpaired t-test with Mann–Whitney test to compare ranks (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, **** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">
18 pages, 3225 KiB  
Article
A Novel Rare PSEN2 Val226Ala in PSEN2 in a Korean Patient with Atypical Alzheimer’s Disease, and the Importance of PSEN2 5th Transmembrane Domain (TM5) in AD Pathogenesis
by YoungSoon Yang, Eva Bagyinszky and Seong Soo A. An
Int. J. Mol. Sci. 2024, 25(17), 9678; https://doi.org/10.3390/ijms25179678 - 6 Sep 2024
Viewed by 848
Abstract
In this manuscript, a novel presenilin-2 (PSEN2) mutation, Val226Ala, was found in a 59-year-old Korean patient who exhibited rapid progressive memory dysfunction and hallucinations six months prior to her first visit to the hospital. Her Magnetic Resonance Imaging (MRI) showed brain atrophy, and [...] Read more.
In this manuscript, a novel presenilin-2 (PSEN2) mutation, Val226Ala, was found in a 59-year-old Korean patient who exhibited rapid progressive memory dysfunction and hallucinations six months prior to her first visit to the hospital. Her Magnetic Resonance Imaging (MRI) showed brain atrophy, and both amyloid positron emission tomography (PET) and multimer detection system-oligomeric amyloid-beta (Aβ) results were positive. The patient was diagnosed with early onset Alzheimer’s disease. The whole-exome analysis revealed a new PSEN2 Val226Ala mutation with heterozygosity in the 5th transmembrane domain of the PSEN2 protein near the lumen region. Analyses of the structural prediction suggested structural changes in the helix, specifically a loss of a hydrogen bond between Val226 and Gln229, which may lead to elevated helix motion. Multiple PSEN2 mutations were reported in PSEN2 transmembrane-5 (TM5), such as Tyr231Cys, Ile235Phe, Ala237Val, Leu238Phe, Leu238Pro, and Met239Thr, highlighting the dynamic importance of the 5th transmembrane domain of PSEN2. Mutations in TM5 may alter the access tunnel of the Aβ substrate in the membrane to the gamma-secretase active site, indicating a possible influence on enzyme function that increases Aβ production. Interestingly, the current patient with the Val226Ala mutation presented with a combination of hallucinations and memory dysfunction. Although the causal mechanisms of hallucinations in AD remain unclear, it is possible that PSEN2 interacts with other disease risk factors, including Notch Receptor 3 (NOTCH3) or Glucosylceramidase Beta-1 (GBA) variants, enhancing the occurrence of hallucinations. In conclusion, the direct or indirect role of PSEN2 Val226Ala in AD onset cannot be ruled out. Full article
(This article belongs to the Special Issue Genetic Research in Neurological Diseases)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Magnetic resonance imaging of the patient: observations of Axial FLAIR (A), (B), (C) sequences of the patient with mild diffuse brain atrophy. (<b>b</b>) Amyloid PET image of the patient: abnormal amyloid deposits observed in gray matter of whole brain, especially in the left temporal lobe. (A) Coronal plane. (B) Axial plane.</p>
Full article ">Figure 2
<p>Sanger sequencing data of patient with PSEN2 Val226Ala mutation.</p>
Full article ">Figure 3
<p>ExPASy predictions for PSEN2 Val226Ala, compared to normal PSEN2 and PSEN2 Val226Ala structure in terms of polarity, Kyte-Doolittle Hydropathy Plots and bulkiness index. The X axis present the residues in PSEN2 (between residue 215 and 227), while the Y axis presents the (<b>a</b>) polarity scores (<b>b</b>) the Kyte-Doolittle Hydropathy Plots (<b>c</b>) and the bulkiness index.</p>
Full article ">Figure 4
<p>(<b>a</b>) Aligned normal and mutant PSEN2 structures. (<b>b</b>) Intramolecular interactions in case of Val226. (<b>c</b>) Intramolecular interactions in case of Ala226. (<b>d</b>) 2D diagram of the intramolecular interaction of Val226 vs. Ala226. The residues which Val226 or Ala226 bind to as covalent bonds are labeled with purple, the hydrogen bonds are labeled with blue, and the Van der Waals bonds are labeled with green.</p>
Full article ">Figure 5
<p>Three-dimensional model of structure of mutations, located in TM5 of PSEN2. (<b>a</b>) Leu225Pro, (<b>b</b>) Glu228Leu, (<b>c</b>) Tyr231Cys, (<b>d</b>) Ile235Phe, (<b>e</b>) Met237Val, (<b>f</b>) Leu238Phe, (<b>g</b>) Leu238Pro, (<b>h</b>) Met239Val, (<b>i</b>) Met239Thr, and (<b>j</b>) Met239Ile.</p>
Full article ">Figure 6
<p>Mutations, located in the 5th transmembrane domain of PSEN2. Variants, which are highlighted in red, were verified to impact amyloid metabolism in cell lines, which are highlighted in red. The variants of which the pathogenic nature remained unclear are highlighted in orange. The location of Val226 is highlighted in yellow.</p>
Full article ">
13 pages, 1159 KiB  
Article
Neuroprotective Effects of Phenolic Constituents from Drynariae Rhizoma
by Jin Sung Ahn, Chung Hyeon Lee, Xiang-Qian Liu, Kwang Woo Hwang, Mi Hyune Oh, So-Young Park and Wan Kyunn Whang
Pharmaceuticals 2024, 17(8), 1061; https://doi.org/10.3390/ph17081061 - 13 Aug 2024
Viewed by 940
Abstract
This study aimed to provide scientific data on the anti-Alzheimer’s disease (AD) effects of phenolic compounds from Drynariae Rhizoma (DR) extract using a multi-component approach. Screening of DR extracts, fractions, and the ten phenolic compounds isolated from DR against the key AD-related enzymes [...] Read more.
This study aimed to provide scientific data on the anti-Alzheimer’s disease (AD) effects of phenolic compounds from Drynariae Rhizoma (DR) extract using a multi-component approach. Screening of DR extracts, fractions, and the ten phenolic compounds isolated from DR against the key AD-related enzymes acetylcholinesterase (AChE), butyrylcholinesterase (BChE), β-site amyloid precursor protein cleaving enzyme 1 (BACE1), and monoamine oxidase-B (MAO-B) confirmed their significant inhibitory activities. The DR extract was confirmed to have BACE1-inhibitory activity, and the ethyl acetate and butanol fractions were found to inhibit all AD-related enzymes, including BACE1, AChE, BChE, and MAO-B. Among the isolated phenolic compounds, compounds (2) caffeic acid 4-O-β-D-glucopyranoside, (6) kaempferol 3-O-rhamnoside 7-O-glucoside, (7) kaempferol 3-o-b-d-glucopyranoside-7-o-a-L-arabinofuranoside, (8) neoeriocitrin, (9) naringin, and (10) hesperidin significantly suppressed AD-related enzymes. Notably, compounds 2 and 8 reduced soluble Amyloid Precursor Protein β (sAPPβ) and β-secretase expression by over 45% at a concentration of 1.0 μM. In the thioflavin T assay, compounds 6 and 7 decreased Aβ aggregation by approximately 40% and 80%, respectively, and degraded preformed Aβ aggregates. This study provides robust evidence regarding the potential of DR as a natural therapeutic agent for AD, highlighting specific compounds that may contribute to its efficacy. Full article
(This article belongs to the Section Natural Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of ten compounds isolated from <span class="html-italic">Drynaria fortunei</span>.</p>
Full article ">Figure 2
<p>The effects of six phenolic compounds (compounds <b>2</b>, <b>6</b>, <b>7</b>, <b>8</b>, <b>9</b>, and <b>10</b>) on sAPPβ and β-secretase production. (<b>A</b>) sAPPβ and β-secretase levels in APP-CHO cells treated with different concentrations (1.0 and 0.5 μM) of six compounds were determined by Western blot analysis. (<b>B</b>,<b>C</b>) Graphs show sAPPβ (<b>B</b>) and β-secretase (<b>C</b>) levels compared to DMSO-treated controls. Values are expressed as a percentage of DMSO-treated control. All data are presented as the mean ± standard deviation of three different experiments. * <span class="html-italic">p</span> &lt; 0.05: significant difference from the DMSO-treated control group.</p>
Full article ">Figure 3
<p>Inhibition of Aβ aggregation and degradation of preformed Aβ aggregates by six phenolic compounds (compounds <b>2</b>, <b>6</b>, <b>7</b>, <b>8</b>, <b>9</b>, and <b>10</b>). (<b>A</b>) Aβ was incubated with six phenolic compounds at concentrations of 50 μM and 10 μM. After 24 h, Aβ aggregation was assessed using the Th T assay. (<b>B</b>) Aβ pre-aggregated for 24 h was exposed to six phenolic compounds at concentrations of 1.0 and 0.5 μM. After another 24 h, Aβ disaggregation was evaluated using the Th T assay. All data are presented as the mean ± standard deviation of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05: significant difference from the Aβ-only group.</p>
Full article ">
18 pages, 5252 KiB  
Article
The Identification of Bioactive Compounds in the Aerial Parts of Agrimonia pilosa and Their Inhibitory Effects on Beta-Amyloid Production and Aggregation
by Chung Hyeon Lee, Min Sung Ko, Ye Seul Kim, Kwang Woo Hwang and So-Young Park
Separations 2024, 11(8), 243; https://doi.org/10.3390/separations11080243 - 9 Aug 2024
Viewed by 1008
Abstract
Alzheimer’s disease (AD) is a progressive neurodegenerative condition characterized by memory and cognitive decline in older individuals. Beta-amyloid (Aβ), a significant component of senile plaques, is recognized as a primary contributor to AD pathology. Hence, substances that can inhibit Aβ [...] Read more.
Alzheimer’s disease (AD) is a progressive neurodegenerative condition characterized by memory and cognitive decline in older individuals. Beta-amyloid (Aβ), a significant component of senile plaques, is recognized as a primary contributor to AD pathology. Hence, substances that can inhibit Aβ production and/or accumulation are crucial for AD prevention and treatment. Agrimonia pilosa LEDEB. (A. pilosa) (Rosaceae), specifically its aerial parts, was identified in our previous screening study as a promising candidate with inhibitory effects on Aβ production. Therefore, in this study, A. pilosa extract was investigated for its anti-amyloidogenic effects, and its bioactive principles were isolated and identified. The ethanol extract of A. pilosa reduced the levels of sAPPβ and β-secretase by approximately 3% and 40%, respectively, compared to the DMSO-treated control group in APP-CHO cells (a cell line expressing amyloid precursor protein), which were similar to those in the positive control group. In addition, the ethanol extract of A. pilosa also hindered Aβ’s aggregation into fibrils and facilitated the disaggregation of Aβ aggregates, as confirmed by a Thioflavin T (Th T) assay. Subsequently, the active constituents were isolated using a bioassay-guided isolation method involving diverse column chromatography. Eleven compounds were identified—epi-catechin (1), catechin (2), (2S, 3S)-dihydrokaempferol 3-O-β-D-glucopyranoside (3), (-)-epiafzelechin 5-O-β-D-glucopyranoside (4), kaempferol 3-O-β-D-glucopyranoside (5), apigenin 7-O-β-D-glucopyranoside (6), dihydrokaempferol 7-O-β-D-glucopyranoside (7), quercetin 3-O-β-D-glucopyranoside (8), (2S, 3S)-taxifolin 3-O-β-D-glucopyranoside (9), luteolin 7-O-β-D-glucopyranoside (10), and apigenin 7-O-β-D-methylglucuronate (11)—identified through 1D and 2D NMR analysis and comparison with data from the literature. These compounds significantly decreased Aβ production by reducing β- and γ-secretase levels. Moreover, none of the compounds affected the expression levels of sAPPα or α-secretase. Further, compounds 1, 2, 4, 8, and 10 demonstrated a dose-dependent reduction in Aβ aggregation and promoted the disaggregation of pre-formed Aβ aggregates. Notably, compound 8 inhibited the aggregation of Aβ into fibrils by about 43% and facilitated the disassembly of Aβ aggregates by 41% compared to the control group containing only Aβ. These findings underscore the potential of A. pilosa extract and its constituents to mitigate a crucial pathological aspect of AD. Therefore, A. pilosa extract and its active constituents hold promise for development as therapeutics and preventatives of AD. Full article
Show Figures

Figure 1

Figure 1
<p>The effect of APE on the levels of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase. (<b>A</b>) The amounts of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase were assessed using Western blot analysis in the APP-CHO cells treated with varying concentrations (100, 50, 25, and 12.5 μg/mL) of APE. The graphs depict the levels of sAPP<span class="html-italic">β</span> (<b>B</b>) and <span class="html-italic">β</span>-secretase (<b>C</b>) relative to those in the DMSO-treated control group. Data are presented as means ± SD of three independent experiments, with the values expressed as a percentage relative to the DMSO-treated control group. * <span class="html-italic">p</span> &lt; 0.05 indicates statistical significance compared to the control group (CTR:DMSO-treated control, PC: positive control, extract of <span class="html-italic">D. crassirhizoma</span> roots, APE: ethanol extract of <span class="html-italic">A. pilosa</span>).</p>
Full article ">Figure 2
<p>The effects of the solvent-partitioned fractions on sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase. (<b>A</b>) The amounts of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase in APP-CHO cells treated with the 4 solvent-partitioned fractions (50 μg/mL) were assessed using Western blot analysis. (<b>B</b>,<b>C</b>) The graphs display the levels of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase, presented as a percentage of the DMSO-treated control group. Data are presented as means ± SD of three independent experiments, with values expressed as a percentage relative to the DMSO-treated control group. * <span class="html-italic">p</span> &lt; 0.05 indicates statistical significance compared to the control group (CTR: DMSO-treated control, PC: positive control, extract of <span class="html-italic">D. crassirhizoma</span> roots, Hx: <span class="html-italic">n</span>-hexane, DCM: dichloromethane, EA: ethyl acetate, DW: water).</p>
Full article ">Figure 3
<p>The effect of APE and its solvent-partitioned fractions on A<span class="html-italic">β</span> aggregation and disaggregation. (<b>A</b>) A<span class="html-italic">β</span> was exposed to concentrations of 100, 20, and 4 μg/mL of APE and the solvent-partitioned fractions. Following a 24 h incubation period, A<span class="html-italic">β</span> aggregation was assessed using the Th T assay. (<b>B</b>) A<span class="html-italic">β</span> aggregates that had been pre-formed for 24 h were exposed to APE and the solvent-partitioned fractions at concentrations of 100, 20, and 4 μg/mL. Following an additional 24 h incubation, the extent of A<span class="html-italic">β</span> aggregation was assessed using the Th T assay. All data are presented as means ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, significantly different from the A<span class="html-italic">β</span>-only group (PC: positive control, quercetin, 50 μg/mL).</p>
Full article ">Figure 4
<p>The structures of the compounds isolated and identified from the EA fraction of APE.</p>
Full article ">Figure 5
<p>HPLC chromatogram of the APE, EA fraction, and isolated compounds. (<b>A</b>) A chromatogram of compounds <b>1</b>–<b>11</b>, (<b>B</b>) APE, and (<b>C</b>) EA fraction recorded at wavelength of 280 nm (APE: <span class="html-italic">A. pliosa</span>, 20 mg/mL; EA: ethyl acetate fraction of <span class="html-italic">A. pliosa</span>, 20 mg/mL; C1: compound <b>1</b>; C2: compound <b>2</b>; C3: compound <b>3</b>; C4: compound <b>4</b>; C5: compound <b>5</b>; C6: compound <b>6</b>; C7: compound <b>7</b>; C8: compound <b>8</b>; C9: compound <b>9</b>; C10: compound <b>10</b>; and C11: compound <b>11</b>; all isolated compounds were analyzed at 1 mg/mL).</p>
Full article ">Figure 6
<p>The effects of compounds <b>1</b>–<b>11</b> on sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase. (<b>A</b>–<b>C</b>,<b>J</b>–<b>L</b>) The supernatant and cell lysates obtained from APP-CHO cells treated with the compounds (50 and 10 μg/mL) were subjected to Western blot analysis to determine the levels of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase, respectively. (<b>D</b>–<b>I</b>,<b>M</b>–<b>R</b>) Graphs showing changes in the levels of sAPP<span class="html-italic">β</span> and <span class="html-italic">β</span>-secretase proteins. Data are presented as means ± SD of three independent experiments, with values expressed as a percentage relative to the DMSO-treated control group. * <span class="html-italic">p</span> &lt; 0.05 indicates statistical significance (PC: positive control, extract of <span class="html-italic">D. crassirhizoma</span> roots).</p>
Full article ">Figure 7
<p>The effects of compounds <b>1</b>–<b>11</b> on the levels of <span class="html-italic">γ</span>-secretase. (<b>A</b>–<b>C</b>, <b>G</b>–<b>I</b>) APP-CHO cells treated with 50 and 10 μg/mL of compounds were subjected to Western blot analysis in order to determine the levels of <span class="html-italic">γ</span>-secretase in the cell lysates. (<b>D</b>–<b>F</b>, <b>J</b>–<b>L</b>) Graphs showing changes in the <span class="html-italic">γ</span>-secretase proteins. Data are presented as means ± SD of three independent experiments, with values expressed as a percentage relative to the DMSO-treated control group. * <span class="html-italic">p</span> &lt; 0.05 indicates statistical significance compared to the control group. * <span class="html-italic">p</span> &lt; 0.05 indicates a significant differences compared to the control group (PC: positive control, extract of <span class="html-italic">D. crassirhizoma</span> roots).</p>
Full article ">Figure 8
<p>The effects of compounds <b>1</b>–<b>11</b> on sAPP<span class="html-italic">α</span> and <span class="html-italic">α</span>-secretase. (<b>A</b>–<b>C</b>,<b>J</b>–<b>L</b>) APP-CHO cells treated with 50 and 10 μg/mL of compounds were subjected to Western blot analysis in order to determine the levels of sAPP<span class="html-italic">α</span> in the supernatant and those of <span class="html-italic">α</span>-secretase in the cell lysates. (<b>D</b>–<b>I</b>,<b>M</b>–<b>R</b>) Graphs showing changes in the levels of sAPP<span class="html-italic">α</span> and <span class="html-italic">α</span>-secretase proteins. Data are presented as means ± SD of three independent experiments, with values expressed as a percentage relative to the DMSO-treated control group. * <span class="html-italic">p</span> &lt; 0.05 indicates statistical significance compared to the control group.</p>
Full article ">Figure 9
<p>The effects of the compounds on the aggregation and disaggregation of A<span class="html-italic">β</span>. (<b>A</b>) The compounds isolated from APE were incubated with A<span class="html-italic">β</span> at concentrations of 100, 50, and 10 μg/mL. After 24 h, the aggregation of A<span class="html-italic">β</span> was assessed using the Th T assay. (<b>B</b>) The compounds isolated from APE were added to A<span class="html-italic">β</span> aggregates pre-formed for 24 h at concentrations of 100, 50, and 10 μg/mL. After an additional 24 h, the disaggregation of A<span class="html-italic">β</span> was evaluated using the Th T assay. All data are presented as means ± SD of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, significantly different from the A<span class="html-italic">β</span>-only group (PC: positive control, quercetin, 50 μg/mL).</p>
Full article ">
12 pages, 1204 KiB  
Communication
Serum Beta-Secretase 1 Activity Is a Potential Marker for the Differential Diagnosis between Alzheimer’s Disease and Frontotemporal Dementia: A Pilot Study
by Claudia Saraceno, Carlo Cervellati, Alessandro Trentini, Daniela Crescenti, Antonio Longobardi, Andrea Geviti, Natale Salvatore Bonfiglio, Sonia Bellini, Roland Nicsanu, Silvia Fostinelli, Gianmarco Mola, Raffaella Riccetti, Davide Vito Moretti, Orazio Zanetti, Giuliano Binetti, Giovanni Zuliani and Roberta Ghidoni
Int. J. Mol. Sci. 2024, 25(15), 8354; https://doi.org/10.3390/ijms25158354 - 30 Jul 2024
Cited by 1 | Viewed by 1066
Abstract
Alzheimer’s disease (AD) and frontotemporal dementia (FTD) are the two major neurodegenerative diseases causing dementia. Due to similar clinical phenotypes, differential diagnosis is challenging without specific biomarkers. Beta-site Amyloid Precursor Protein cleaving enzyme 1 (BACE1) is a β-secretase pivotal in AD pathogenesis. In [...] Read more.
Alzheimer’s disease (AD) and frontotemporal dementia (FTD) are the two major neurodegenerative diseases causing dementia. Due to similar clinical phenotypes, differential diagnosis is challenging without specific biomarkers. Beta-site Amyloid Precursor Protein cleaving enzyme 1 (BACE1) is a β-secretase pivotal in AD pathogenesis. In AD and mild cognitive impairment subjects, BACE1 activity is increased in brain/cerebrospinal fluid, and plasma levels appear to reflect those in the brain. In this study, we aim to evaluate serum BACE1 activity in FTD, since, to date, there is no evidence about its role. The serum of 30 FTD patients and 30 controls was analyzed to evaluate (i) BACE1 activity, using a fluorescent assay, and (ii) Glial Fibrillary Acid Protein (GFAP) and Neurofilament Light chain (NfL) levels, using a Simoa kit. As expected, a significant increase in GFAP and NfL levels was observed in FTD patients compared to controls. Serum BACE1 activity was not altered in FTD patients. A significant increase in serum BACE1 activity was shown in AD vs. FTD and controls. Our results support the hypothesis that serum BACE1 activity is a potential biomarker for the differential diagnosis between AD and FTD. Full article
Show Figures

Figure 1

Figure 1
<p>BACE1 activity (<b>a</b>) and the levels of GFAP (<b>b</b>) and NfL (<b>c</b>) in the serum of CTRL (<span class="html-italic">n</span> = 30) and FTD patients (<span class="html-italic">n</span> = 30). No differences were observed in BACE1 activity between the two groups. A significant increase in GFAP and NfL levels was shown in FTD patients compared to CTRL. Mean ± SEM; * <span class="html-italic">p</span> &lt; 0.05 and **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>BACE1 activity (<b>a</b>) and the levels of GFAP (<b>b</b>) and NfL (<b>c</b>) in the serum of CTRL (<span class="html-italic">n</span> = 60), AD (<span class="html-italic">n</span> = 31), and FTD (<span class="html-italic">n</span> = 30) patients. A significant increase in BACE1 activity was shown in AD patients compared to CTRL and FTD patients. No differences were observed between CTRL and FTD patients. A significant increase in both GFAP and NfL levels was observed in AD and FTD patients compared to CTRL. Moreover, a significant increase in GFAP levels was shown in AD compared to FTD patients. No differences of NfL levels were observed between AD and FTD patients. Mean ± SEM; * <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>Classification tree obtained for AD (<span class="html-italic">n</span> = 31) vs. FTD (<span class="html-italic">n</span> = 30) patients based on BACE1 activity and GFAP levels in serum. BACE1 activity resulted in being able to discriminate AD from FTD patients with very high percentage (95.7%).</p>
Full article ">
17 pages, 4197 KiB  
Article
Conformational Changes and Unfolding of β-Amyloid Substrates in the Active Site of γ-Secretase
by Jakub Jakowiecki, Urszula Orzeł, Przemysław Miszta, Krzysztof Młynarczyk and Sławomir Filipek
Int. J. Mol. Sci. 2024, 25(5), 2564; https://doi.org/10.3390/ijms25052564 - 22 Feb 2024
Viewed by 1215
Abstract
Alzheimer’s disease (AD) is the leading cause of dementia and is characterized by a presence of amyloid plaques, composed mostly of the amyloid-β (Aβ) peptides, in the brains of AD patients. The peptides are generated from the amyloid precursor protein (APP), which undergoes [...] Read more.
Alzheimer’s disease (AD) is the leading cause of dementia and is characterized by a presence of amyloid plaques, composed mostly of the amyloid-β (Aβ) peptides, in the brains of AD patients. The peptides are generated from the amyloid precursor protein (APP), which undergoes a sequence of cleavages, referred as trimming, performed by γ-secretase. Here, we investigated conformational changes in a series of β-amyloid substrates (from less and more amyloidogenic pathways) in the active site of presenilin-1, the catalytic subunit of γ-secretase. The substrates are trimmed every three residues, finally leading to Aβ40 and Aβ42, which are the major components of amyloid plaques. To study conformational changes, we employed all-atom molecular dynamics simulations, while for unfolding, we used steered molecular dynamics simulations in an implicit membrane-water environment to accelerate changes. We have found substantial differences in the flexibility of extended C-terminal parts between more and less amyloidogenic pathway substrates. We also propose that the positively charged residues of presenilin-1 may facilitate the stretching and unfolding of substrates. The calculated forces and work/energy of pulling were exceptionally high for Aβ40, indicating why trimming of this substrate is so infrequent. Full article
(This article belongs to the Section Molecular Biophysics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A comparison of cleavage products of amyloid precursor protein (APP). APP-C83 is generated by α-secretase cleavage, while APP-C99 is generated by β-secretase cleavage. These pathways are called nonamyloidogenic and amyloidogenic, respectively. Below are four intermediate products obtained from APP-C99, which are substrates for subsequent cuts by γ-secretase: Aβ<sub>48</sub> and Aβ<sub>45</sub> leading to Aβ<sub>42</sub> (more amyloidogenic product), and Aβ<sub>46</sub> and Aβ<sub>43</sub> leading to Aβ<sub>40</sub> (less amyloidogenic product). Black rectangles denote sequences used to create substrates of the same length for MD simulations.</p>
Full article ">Figure 2
<p>The structure of γ-secretase (GS). (<b>a</b>) The cryo-EM structure (PDB id:6IYC) of GS with APP-C83 substrate. Colors of subunits: PS-1 in cyan, APH-1 in purple, NCT in green, PEN-2 in yellow, and the Aβ substrate in salmon. (<b>b</b>) Magnification of the catalytic site showing β-sheet formed between a substrate and PS-1. The catalytic residues of PS-1 are shown in red. (<b>c</b>) The substrate after threading of substrate sequence into 6IYC structure to obtain the substrate extended by three residues behind the catalytic residues for trimming.</p>
Full article ">Figure 3
<p>Two-dimensional scatter plots showing distances between the carboxylic group of one of the catalytic residues, Asp385, to the peptide bond of residue that is to be cleaved (horizontal axes) and distances between carboxylic groups of the catalytic residues (vertical axes). Each panel is averaged from two MD simulations. All points are colored according to the MD simulation time from purple (0 ns) to yellow (500 ns). (<b>a</b>) Aβ<sub>46</sub> substrate with a Thr43 scissile bond; (<b>b</b>) Aβ<sub>43</sub> substrate with a Val40 scissile bond; (<b>c</b>) Aβ<sub>48</sub> substrate with an Ile45 scissile bond; (<b>d</b>) Aβ<sub>45</sub> substrate with a Ala42 scissile bond.</p>
Full article ">Figure 4
<p>Two-dimensional scatter plots showing distances between C<sub>α</sub> (CA) atoms of selected residues from the substrate and PS-1: the distance between C<sub>α</sub> atoms of the last residue (n) of the substrate and Lys380 of PS-1 (horizontal axes), and the distance between C<sub>α</sub> atoms of last but one residue (n − 1) of the substrate and Leu381 of PS-1 (vertical axes). Each panel is averaged from two MD simulations. Red dashed square indicates the shortest distances between the above residues with a possibility of forming a β-sheet composed of four residues: n and n-1 substrate residues and PS-1 residues 380–381. (<b>a</b>) Aβ<sub>46</sub> substrate; (<b>b</b>) Aβ<sub>43</sub> substrate; (<b>c</b>) Aβ<sub>48</sub> substrate; (<b>d</b>) Aβ<sub>45</sub> substrate.</p>
Full article ">Figure 5
<p>Timeline showing the secondary structure of four substrates during 500 ns of MD simulation. Each simulation is presented in a separate graph over time along the horizontal axis. N-terminal part of substrates is not shown. (<b>a</b>) Aβ<sub>46</sub> substrate; (<b>b</b>) Aβ<sub>43</sub> substrate; (<b>c</b>) Aβ<sub>48</sub> substrate; (<b>d</b>) Aβ<sub>45</sub> substrate.</p>
Full article ">Figure 6
<p>Lys380 residue forms a salt bridge with the substrate C-terminus; (<b>a</b>) with Thr43 of Aβ<sub>43</sub>; (<b>b</b>) with Ile45 of Aβ<sub>45</sub>; (<b>c</b>) with Val46 of Aβ<sub>46</sub>; (<b>d</b>) with Thr48 of Aβ<sub>48</sub>; (<b>e</b>) 3D structure of Aβ<sub>48</sub> substrate extended by three residues so the catalytic residues (in red) are localized in proximity of next cut residue Ile45. Red dashed rectangles in panels (<b>a</b>–<b>c</b>) indicate formation of a salt bridge. Results from two independent simulations in panels (<b>a</b>–<b>d</b>) are shown in different colors. The hydrogen bonds are indicated by dashed yellow cylinders while distances are shown in [Å]. The catalytic residues are shown in red, PS-1 in cyan and Aβ in salmon.</p>
Full article ">Figure 7
<p>Two arginine residues, Arg269 and Arg377, of PS-1 form a salt bridge with C-terminus of the substrate. (<b>a</b>) The distance between the carboxyl terminal group of C-terminus of Aβ<sub>48</sub> (atom O) and Arg269 of PS-1 (atom CZ). (<b>b</b>) The distance between the carboxyl terminal group of C-terminus of Aβ<sub>48</sub> (atom O) and Arg377 of PS-1 (atom CZ). Results from two simulations are shown in different colors. Red dashed rectangles indicate formation of a salt bridge and a hydrogen bond. (<b>c</b>) Magnification of the area showing C-terminal residues of Aβ<sub>48</sub> substrate (in salmon) and residues of PS-1 (in cyan): two arginine residues, Arg269 and Arg377, in orange and two catalytic residues, Asp257 and Asp385 (in red). The hydrogen bonds are indicated by dashed yellow cylinders while distances are shown in [Å]. PS-1 is shown in cyan and Aβ in salmon.</p>
Full article ">Figure 8
<p>Final representative conformations of substrates after SMD simulations in GS-SMD web server. Unfolding was performed until the conformation ready for next trimming was reached (by three residues). GS is not shown so as not to obscure the substrate. (<b>a</b>) Pulling and unfolding of Aβ<sub>49</sub> to prepare the next cleavage event Aβ<sub>49</sub> → Aβ<sub>46</sub>; (<b>b</b>) unfolding of Aβ<sub>46</sub> → Aβ<sub>43</sub>; (<b>c</b>) unfolding of Aβ<sub>43</sub> → Aβ<sub>40</sub>; (<b>d</b>) unfolding of Aβ<sub>40</sub> → Aβ<sub>37</sub>. Borders of the hydrophobic core of the membrane are represented by two planes. The polar and charged residues remain on the extracellular side and pull the substrate back. Colors of residues: green—apolar, blue—polar, red—charged.</p>
Full article ">Figure 9
<p>Box-plot representing the work/energy required to unfold the substrate by three residues in the active site of GS during SMD simulations. For each substrate, eight independent simulations were conducted. The work/energy was calculated for that frame with the shortest sum of distances of the n − 3 residue to the catalytic residues.</p>
Full article ">Figure 10
<p>GS-SMD simulations—possible cleavage event. A frame with the shortest distance between GS catalytic residues (Asp257 and protonated Asp385) and the n − 3 substrate residue (Thr30) in the simulation of substrate unfolding Aβ<sub>49</sub> → Aβ<sub>46</sub>. The sequence number of the C-terminal residue in the GS-SMD server is always 33, regardless of the thread sequence. Color scheme: PS-1 in cyan and Aβ in salmon.</p>
Full article ">
10 pages, 2594 KiB  
Case Report
PSEN1 His214Asn Mutation in a Korean Patient with Familial EOAD and the Importance of Histidine–Tryptophan Interactions in TM-4 Stability
by Eva Bagyinszky, Minju Kim, Young Ho Park, Seong Soo A. An and SangYun Kim
Int. J. Mol. Sci. 2024, 25(1), 116; https://doi.org/10.3390/ijms25010116 - 21 Dec 2023
Cited by 1 | Viewed by 1101
Abstract
A pathogenic mutation in presenilin-1 (PSEN1), His214Asn, was found in a male patient with memory decline at the age of 41 in Korea for the first time. The proband patient was associated with a positive family history from his father, paternal [...] Read more.
A pathogenic mutation in presenilin-1 (PSEN1), His214Asn, was found in a male patient with memory decline at the age of 41 in Korea for the first time. The proband patient was associated with a positive family history from his father, paternal aunt, and paternal grandmother without genetic testing. He was diagnosed with early onset Alzheimer’s disease (EOAD). PSEN1 His214Asn was initially reported in an Italian family, where the patient developed phenotypes similar to the current proband patient. Magnetic resonance imaging (MRI) scans revealed a mild hippocampal atrophy. The amyloid positron emission tomography (amyloid-PET) was positive, along with the positive test results of the increased amyloid ß (Aβ) oligomerization tendency with blood. The PSEN1 His214 amino acid position plays a significant role in the gamma–secretase function, especially from three additional reported mutations in this residue: His214Asp, His214Tyr, and His214Arg. The structure prediction model revealed that PSEN1 protein His214 may interact with Trp215 of His-Trp cation-π interaction, and the mutations of His214 would destroy this interaction. The His-Trp cation-π interaction between His214 and Trp215 would play a crucial structural role in stabilizing the 4th transmembrane domain of PSEN1 protein, especially when aromatic residues were often reported in the membrane interface of the lipid–extracellular region of alpha helices or beta sheets. The His214Asn would alter the cleavage dynamics of gamma–secretase from the disappeared interactions between His214 and Trp215 inside of the helix, resulting in elevated amyloid production. Hence, the increased Aβ was reflected in the increased Aβ oligomerization tendency and the accumulations of Aβ in the brain from amyloid-PET, leading to EOAD. Full article
(This article belongs to the Special Issue The Role of Genetics in Dementia)
Show Figures

Figure 1

Figure 1
<p>Family tree of proband patient with <span class="html-italic">PSEN1</span> His214Asn (III-1). Three additional family members (I-2, II-1 and II-4) also developed memory dysfunctions, but segregation could not be proven. Black colors are the family members, affected with memory decline. Black arrow shows the proband patient.</p>
Full article ">Figure 2
<p>Neuroimaging data for the proband (III-1). (<b>A</b>). Magnetic resonance imaging (MRI) revealed mild hippocampal atrophy. (<b>B</b>). MRI showed a mild white matter change. (<b>C</b>). Amyloid-positron emission tomography (PET) indicated amyloid positivity in several brain regions (bilateral– lateral temporal, frontal, and parietal areas).</p>
Full article ">Figure 3
<p>Sanger sequencing of patient with heterozygous <span class="html-italic">PSEN1</span> His214Asn. Red arrow shows the mutation.</p>
Full article ">Figure 4
<p>Structure predictions of (<b>a</b>) PSEN1 protein (helix-4) with His214 and (<b>b</b>) PSEN1 with Asn214. Even though the interactions between Asn214 to both Ile211 and Met210 remained in both structures, a cation-π interaction between Asn214 and Trp215 disappeared, suggesting the maintenance of helix within plasma membrane. The position of the helix would be affected, altering the γ-sec functions.</p>
Full article ">Figure 5
<p>STRING analysis of rare variant-carrying genes found in the proband patient. Mutations in PSEN1 would affect its interactions with other correlating genes—LRRK2, NME8, NOTCH3, and ZCWPW1 directly and RIN3 indirectly. The green lines show that these proteins may interact based on “textmining”, and there is proof of interaction, based on literature. The black lines mean the genes may have evidence for co-expression. The blue line shows the gene interaction has been proven, based on curated databases (such as Biocarta or Reactome). The purple line shows the interaction had been proven experimentally. Length of lines are customizable.</p>
Full article ">Figure 6
<p>Location of His214 in the 4th transmembrane domain of PSEN1 protein. The position of the Asn214 mutation would affect the position of the helix in helix–helix interactions within plasma membrane and the dynamics of the γ-sec functions.</p>
Full article ">
17 pages, 6537 KiB  
Article
Neuroprotective Effects of N-methyl-(2S, 4R)-trans-4-hydroxy-L-proline (NMP) against Amyloid-β-Induced Alzheimer’s Disease Mouse Model
by Jawad Ali, Amjad Khan, Jun Sung Park, Muhammad Tahir, Waqas Ahmad, Kyonghwan Choe and Myeong Ok Kim
Nutrients 2023, 15(23), 4986; https://doi.org/10.3390/nu15234986 - 1 Dec 2023
Cited by 5 | Viewed by 2352
Abstract
Alzheimer’s disease (AD), is a progressive neurodegenerative disorder that involves the deposition of β-amyloid plaques and the clinical symptoms of confusion, memory loss, and cognitive dysfunction. Despite enormous progress in the field, no curative treatment is available. Therefore, the current study was designed [...] Read more.
Alzheimer’s disease (AD), is a progressive neurodegenerative disorder that involves the deposition of β-amyloid plaques and the clinical symptoms of confusion, memory loss, and cognitive dysfunction. Despite enormous progress in the field, no curative treatment is available. Therefore, the current study was designed to determine the neuroprotective effects of N-methyl-(2S, 4R)-Trans-4-hydroxy-L-proline (NMP) obtained from Sideroxylon obtusifolium, a Brazilian folk medicine with anti-inflammatory and anti-oxidative properties. Here, for the first time, we explored the neuroprotective role of NMP in the Aβ1–42-injected mouse model of AD. After acclimatization, a single intracerebroventricular injection of Aβ1–42 (5 µL/5 min/mouse) in C57BL/6N mice induced significant amyloidogenesis, reactive gliosis, oxidative stress, neuroinflammation, and synaptic and memory deficits. However, an intraperitoneal injection of NMP at a dose of (50 mg/kg/day) for three consecutive weeks remarkably decreased beta secretase1 (BACE-1) and Aβ, activated the astrocyte and microglia expression level as well as downstream inflammatory mediators such as pNF-ĸB, TNF-α, and IL-1β. NPM also strongly attenuated oxidative stress, as evaluated by the expression level of NRF2/HO-1, and synaptic failure, by improving the level of both the presynaptic (SNAP-25 and SYN) and postsynaptic (PSD-95 and SNAP-23) regions of the synapses in the cortexes and hippocampi of the Aβ1–42-injected mice, contributing to cognitive improvement in AD and improving the behavioral deficits displayed in the Morris water maze and Y-maze. Overall, our data suggest that NMP provides potent multifactorial effects, including the inhibition of amyloid plaques, oxidative stress, neuroinflammation, and cognitive deficits. Full article
(This article belongs to the Special Issue Oxidative Stress and Protective Effects of Natural Products in Health)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Experimental plan for the current study of NMP against Aβ<sub>1–42</sub>-induced AD mice model. (<b>B</b>) Chemical structure of N-methyl-(2S, 4R)-Trans-4-hydroxy-L-proline.</p>
Full article ">Figure 2
<p>The impact of NMP on Alzheimer’s protein indicators within the brains of mice induced with Aβ<sub>1–42</sub>. (<b>A</b>) Immunoblot analyses and bar graphs depicting the levels of BACE-1 and Aβ expression in the cortex and hippocampus of mouse brains after the administration of Aβ<sub>1–42</sub> and NMP. (<b>B</b>) Representative images and a corresponding bar graph showing the relative integrated density for Aβ in the cortical and hippocampal tissue (DG region) of mouse brains (<span class="html-italic">n</span> = 4 mice/group). Photomicrograph of (10X) magnification and inset scale bar is 50 µm. Band intensities were measured using ImageJ software (ver. 8.0, San Diego, CA, USA), and the distinctions were illustrated through a bar graph generated by GraphPad Prism. Beta-actin was utilized as a reference for loading. The mean ± S.E.M values for the indicated proteins are displayed as relative integrated density levels (<span class="html-italic">n</span> = 4 mice/group). ** <span class="html-italic">p</span> &lt; 0.01 signify a notable distinction compared to the Aβ<sub>1–42</sub>-treated group, while ## <span class="html-italic">p</span> &lt; 0.01 signifies a significant contrast from the vehicle-treated group.</p>
Full article ">Figure 3
<p>Effects of NMP on astrocytosis and microgliosis within the brains of mice injected with Aβ<sub>1–42</sub>. (<b>A</b>) Images of the scanned Western blot results and bar graph for indicated (GFAP and Iba-1) protein expression in the cortex and hippocampus of mice brain following Aβ<sub>1–42</sub> and NMP treatment. The differences are shown in the bar graph. (<b>B</b>) Illustrative images along with an associated bar graph displaying relative integrated density of GFAP in the cortex and hippocampus (DG region) of mouse brains. Photomicrograph of (10X) magnification and inset scale bar is 50 µm. The information is displayed as the average value ± standard error of the mean (<span class="html-italic">n</span> = 4 mice per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Aβ<sub>1–42</sub>-treated group and ## <span class="html-italic">p</span> &lt; 0.01 vs. vehicle-treated group.</p>
Full article ">Figure 4
<p>Effects of NMP on oxidative stress in the brain of Aβ<sub>1–42</sub>-induced mice. (<b>A</b>) Immunoblot analyses and bar graphs depicting the protein expression levels of NRF2 and HO-1 in the cortex and hippocampus of mouse brains after Aβ<sub>1–42</sub> and NMP treatment. (<b>B</b>) Representative photographs and a corresponding bar graph showing relative integrated density of NRF2 in the cortex and hippocampus (DG region) of mouse brains. Photomicrograph of 10X magnification, and inset scale bar is 50 µm. Data are presented as the mean ± S.E.M (<span class="html-italic">n</span> = 4 mice/group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Aβ<sub>1–42</sub>-treated group and ## <span class="html-italic">p</span> &lt; 0.01 vs. vehicle-treated group.</p>
Full article ">Figure 5
<p>Effects of NMP on inflammatory cytokines as well as apoptotic marker in the brain of Aβ<sub>1–42</sub>-induced mice. The Western blot assessment and graphical representations indicating the protein expression levels of (pNF-κB, TNF-α, and IL-1β) in the cortexes and hippocampi of mouse brains after Aβ<sub>1–42</sub> and NMP treatment. The bands were assessed and measured utilizing ImageJ software, and the differences are displayed in the bar chart. β-actin served as the standard for loading. The levels of relative density are presented in arbitrary units (A.U.) as the mean ± S.E.M for the specified proteins (<span class="html-italic">n</span> = 4 mice per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Aβ<sub>1–42</sub>-treated group and ## <span class="html-italic">p</span> &lt; 0.01 vs. vehicle-treated group.</p>
Full article ">Figure 6
<p>Effects of NMP on synaptic proteins in the brain of Aβ<sub>1–42</sub>-induced mice. Immunoblot analysis and bar graphs for the cortical and hippocampal protein expression levels of (PSD-95, SNAP-25, SNAP-23, and Synaptophysin) in the brains of mice, followed by Aβ<sub>1–42</sub> and NMP administration. ImageJ software was used for the determination of band densities of these synaptic markers, while the differences are represented by the bar graph that was produced with GraphPad Prism 8 software. The data is displayed as the average ± standard error of the mean (<span class="html-italic">n</span> = 4 mice per group). ** <span class="html-italic">p</span> &lt; 0.01 vs. Aβ<sub>1–42</sub>-treated group and ## <span class="html-italic">p</span> &lt; 0.01 vs. vehicle-treated group. β-actin was used as a loading standard.</p>
Full article ">Figure 7
<p>Effects of NMP on memory impairment and cognitive dysfunction in Aβ<sub>1–42</sub>-induced mice. (<b>A</b>) Images of the trajectory map in the MWM and Y-maze task. (<b>B</b>) Line graph showing mean escape latency during training days to reach the visible platform in the MWM task. (<b>C</b>) Time spent in the designated quadrant during the probe trial. (<b>D</b>) Number of crossings around platform during the probe trial. (<b>E</b>) Y-maze task for the measurement of spontaneous alteration behavior percentage in respective groups. The results are shown as the mean ± SEM (<span class="html-italic">n</span> = 8 mice/group). * <span class="html-italic">p</span> &lt; 0.05, and ** <span class="html-italic">p</span> &lt; 0.01 vs. Aβ<sub>1–42</sub>-treated group and # <span class="html-italic">p</span> &lt; 0.05, and ## <span class="html-italic">p</span> &lt; 0.01 vs. vehicle-treated group.</p>
Full article ">Figure 8
<p>Graphical abstract showing the possible neuroprotective effects of NMP in Aβ<sub>1–42</sub>-induced mice. An accumulation of amyloid beta (Aβ<sub>1–42</sub>) stimulates amyloidogenesis, reactive gliosis, oxidative stress neuroinflammation, and synaptic and memory deficits. These effects are mitigated by NMP in neurodegenerative disorders by reducing the burden of amyloid plaques by reducing the amyloid plaques, gliosis, oxidative stress, as well as neuroinflammation by decreasing the expression level of BACE-1, Aβ, GFAP, Iba-1, ROS, and inflammatory cytokines, while increasing cognitive function by regulating synaptic markers.</p>
Full article ">
27 pages, 6218 KiB  
Article
Development of Potential Multi-Target Inhibitors for Human Cholinesterases and Beta-Secretase 1: A Computational Approach
by Deyse B. Barbosa, Mayra R. do Bomfim, Tiago A. de Oliveira, Alisson M. da Silva, Alex G. Taranto, Jorddy N. Cruz, Paulo B. de Carvalho, Joaquín M. Campos, Cleydson B. R. Santos and Franco H. A. Leite
Pharmaceuticals 2023, 16(12), 1657; https://doi.org/10.3390/ph16121657 - 28 Nov 2023
Cited by 3 | Viewed by 1557
Abstract
Alzheimer’s disease causes chronic neurodegeneration and is the leading cause of dementia in the world. The causes of this disease are not fully understood but seem to involve two essential cerebral pathways: cholinergic and amyloid. The simultaneous inhibition of AChE, BuChE, and BACE-1, [...] Read more.
Alzheimer’s disease causes chronic neurodegeneration and is the leading cause of dementia in the world. The causes of this disease are not fully understood but seem to involve two essential cerebral pathways: cholinergic and amyloid. The simultaneous inhibition of AChE, BuChE, and BACE-1, essential enzymes involved in those pathways, is a promising therapeutic approach to treat the symptoms and, hopefully, also halt the disease progression. This study sought to identify triple enzymatic inhibitors based on stereo-electronic requirements deduced from molecular modeling of AChE, BuChE, and BACE-1 active sites. A pharmacophore model was built, displaying four hydrophobic centers, three hydrogen bond acceptors, and one positively charged nitrogen, and used to prioritize molecules found in virtual libraries. Compounds showing adequate overlapping rates with the pharmacophore were subjected to molecular docking against the three enzymes and those with an adequate docking score (n = 12) were evaluated for physicochemical and toxicological parameters and commercial availability. The structure exhibiting the greatest inhibitory potential against all three enzymes was subjected to molecular dynamics simulations (100 ns) to assess the stability of the inhibitor-enzyme systems. The results of this in silico approach indicate ZINC1733 can be a potential multi-target inhibitor of AChE, BuChE, and BACE-1, and future enzymatic assays are planned to validate those results. Full article
Show Figures

Figure 1

Figure 1
<p>ROC curves obtained for pharmacophore models of AChE, BuChE, and BACE-1 inhibitors (A = AUC ROC).</p>
Full article ">Figure 2
<p>Representation of the best pharmacophore model for AChE, BuChE, and BACE-1 inhibitors (cyan spheres: hydrophobic centers; green: H-bond acceptors; red: positively charged center). The size of the spheres varies according to the tolerance radius calculated by GALAHAD<sup>TM</sup>. Numbers represent distances in Angstroms.</p>
Full article ">Figure 3
<p>Structures selected through hierarchical virtual screening: (<b>a</b>) ZINC1733, (<b>b</b>) ZINC6063.</p>
Full article ">Figure 4
<p>Representation of the intermolecular interactions of the crystallographic ligand dihydrotanshinone I (<b>a</b>), ZINC1733 (<b>b</b>) and ZINC6063 (<b>c</b>) at the AChE binding site.</p>
Full article ">Figure 5
<p>Representation of the intermolecular interactions of the crystallographic ligand tacrine (<b>a</b>), ZINC1733 (<b>b</b>) and ZINC6063 (<b>c</b>) at the BuChE binding site.</p>
Full article ">Figure 6
<p>Representation of the intermolecular interactions of the crystallographic ligand 6UWP (<b>a</b>), ZINC1733 (<b>b</b>) and ZINC6063 (<b>c</b>) at the BACE-1 binding site.</p>
Full article ">Figure 7
<p>RMSD (backbone) of APO structures and AChE (<b>a</b>), BuChE (<b>b</b>), and BACE-1 (<b>c</b>) complexes with ZINC1733 during the 100 ns molecular dynamics.</p>
Full article ">Figure 8
<p>RMSF (Å) (backbone) of APO structures and complexes of AChE (<b>a</b>), BuChE (<b>b</b>), and BACE-1 (<b>c</b>) with ZINC1733 during the respective productive phase. The blue highlights correspond to the binding site regions.</p>
Full article ">Figure 9
<p>Permanence rate of hydrogen interactions (hbond) of ZINC1733 in the active site of AChE, BuChE and BACE-1 during the production phase and identification of the involved pairs.</p>
Full article ">Figure 10
<p>Interactions of ZINC1733 at the AChE binding site obtained from the MD simulation representative structure (69.5 ns).</p>
Full article ">Figure 11
<p>Interactions of ZINC1733 at the BuChE binding site obtained from the MD simulation representative structure (45.70 ns).</p>
Full article ">Figure 12
<p>Interactions of ZINC1733 at the BACE-1 binding site obtained from the MD simulation representative structure (90.60 ns).</p>
Full article ">
Back to TopTop