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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,148)

Search Parameters:
Keywords = amyloid β-protein

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 3616 KiB  
Article
A Structural Effect of the Antioxidant Curcuminoids on the Aβ(1–42) Amyloid Peptide
by Angelo Santoro, Antonio Ricci, Manuela Rodriquez, Michela Buonocore and Anna Maria D’Ursi
Antioxidants 2025, 14(1), 53; https://doi.org/10.3390/antiox14010053 (registering DOI) - 5 Jan 2025
Abstract
Investigating amyloid–β (Aβ) peptides in solution is essential during the initial stages of developing lead compounds that can influence Aβ fibrillation while the peptide is still in a soluble state. The tendency of the Aβ(1–42) peptide to misfold in solution, correlated to the [...] Read more.
Investigating amyloid–β (Aβ) peptides in solution is essential during the initial stages of developing lead compounds that can influence Aβ fibrillation while the peptide is still in a soluble state. The tendency of the Aβ(1–42) peptide to misfold in solution, correlated to the aetiology of Alzheimer’s disease (AD), is one of the main hindrances to characterising its aggregation kinetics in a cell-mimetic environment. Moreover, the Aβ(1–42) aggregation triggers the unfolded protein response (UPR) in the endoplasmic reticulum (ER), leading to cellular dysfunction and multiple cell death modalities, exacerbated by reactive oxygen species (ROS), which damage cellular components and trigger inflammation. Antioxidants like curcumin, a derivative of Curcuma longa, help mitigate ER stress by scavenging ROS and enhancing antioxidant enzymes. Furthermore, evidence in the literature highlights the effect of curcumin on the secondary structure of Aβ(1–42). This explorative study investigates the Aβ(1–42) peptide conformational behaviour in the presence of curcumin and six derivatives using circular dichroism (CD) to explore their interactions with lipid bilayers, potentially preventing aggregate formation. The results suggest that the synthetic tetrahydrocurcumin (THC) derivative interacts with the amyloid peptide in all the systems presented, while cyclocurcumin (CYC) and bisdemethoxycurcumin (BMDC) only interact when the peptide is in a less stable conformation. Molecular dynamics simulations helped visualise the curcuminoids’ effect in an aqueous system and hypothesise the importance of the peptide surface exposition to the solvent, differently modulated by the curcumin derivatives. Full article
Show Figures

Figure 1

Figure 1
<p>CD spectra and secondary structure quantification performed with CONTIN algorithm of Aβ(1–42) free peptide and in the presence of curcumin derivatives in DOPC SUVs. The results of the statistical analysis conducted using the one-way ANOVA method are reported in the table.</p>
Full article ">Figure 2
<p>CD spectra and secondary structure quantification performed with CONTIN algorithm of Aβ(1–42) free peptide and in the presence of curcumin derivatives in DOPG liposomal system. The results of the statistical analysis conducted using the one-way ANOVA method are reported in the table.</p>
Full article ">Figure 3
<p>CD spectra and secondary structure quantification performed with CONTIN algorithm of Aβ(1–42) free peptide and in the presence of curcumin derivatives in HFIP/H<sub>2</sub>O 80/20 <span class="html-italic">v</span>/<span class="html-italic">v</span>. The results of the statistical analysis conducted using the one-way ANOVA method are reported in the table.</p>
Full article ">Figure 4
<p>CD spectra and secondary structure quantification performed with CONTIN algorithm of Aβ(1–42) free peptide and in the presence of curcumin derivatives in HFIP/H<sub>2</sub>O 50/50 <span class="html-italic">v</span>/<span class="html-italic">v</span>. The results of the statistical analysis conducted using the one-way ANOVA method are reported in the table.</p>
Full article ">Figure 5
<p>CD spectra and secondary structure quantification performed with CONTIN algorithm of Aβ(1–42) free peptide and in the presence of curcumin derivatives in HFIP/H<sub>2</sub>O 20/80 <span class="html-italic">v</span>/<span class="html-italic">v</span>. The results of the statistical analysis conducted using the one-way ANOVA method are reported in the table.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>C</b>) Snapshots of the last steps of the three independent 100 ns MD simulations performed on five Aβ(1–42) monomers in water at pH 7.4. The peptides are reported in ribbon representation and coloured according to the chain name (A–E, red to purple). The α-helix moieties are depicted as wider spiral ribbons, while the β-sheets are indicated as flat arrowed ribbons. (<b>D</b>) Selected structures of Aβ(1–42) extracted from the MD simulations compared to (<b>E</b>) structures of Aβ(1–42) in the bundle obtained by acquiring 2D NMR spectra in HFIP/H<sub>2</sub>O 50/50 <span class="html-italic">v</span>/<span class="html-italic">v</span>; the ribbons here are coloured according to the residue position in the peptide (N-terminus to C-terminus, red to purple).</p>
Full article ">Figure 7
<p>MD results of Aβ(1–42) in complex with curcuminoids (BDMC, CUR, CYC, OHC, THC). Secondary structure prediction of the five amyloid monomers in the presence of the curcuminoids (<b>A</b>) in the “unbiased simulation” and (<b>B</b>) in the “biased simulation”. (<b>C</b>) Plot of the solvent-accessible surface area of all peptides in the presence of the curcuminoids as a function of the simulation time from the “unbiased simulation”. (<b>D</b>) Heatmap of the solvent-accessible surface area for each of the five Aβ(1–42) monomers (chains A–E) in the presence of the curcuminoids as a function of the simulation time from the “unbiased simulation”. The colour code in the bar below represents the solvent exposure, ranging from the smallest area measured (34 Å<sup>2</sup>) to the largest (60 Å<sup>2</sup>).</p>
Full article ">
90 pages, 16171 KiB  
Perspective
Production of Amyloid-β in the Aβ-Protein-Precursor Proteolytic Pathway Is Discontinued or Severely Suppressed in Alzheimer’s Disease-Affected Neurons: Contesting the ‘Obvious’
by Vladimir Volloch and Sophia Rits-Volloch
Genes 2025, 16(1), 46; https://doi.org/10.3390/genes16010046 - 2 Jan 2025
Viewed by 172
Abstract
A notion of the continuous production of amyloid-β (Aβ) via the proteolysis of Aβ-protein-precursor (AβPP) in Alzheimer’s disease (AD)-affected neurons constitutes both a cornerstone and an article of faith in the Alzheimer’s research field. The present Perspective challenges this assumption. It analyses the [...] Read more.
A notion of the continuous production of amyloid-β (Aβ) via the proteolysis of Aβ-protein-precursor (AβPP) in Alzheimer’s disease (AD)-affected neurons constitutes both a cornerstone and an article of faith in the Alzheimer’s research field. The present Perspective challenges this assumption. It analyses the relevant empirical data and reaches an unexpected conclusion, namely that in AD-afflicted neurons, the production of AβPP-derived Aβ is either discontinued or severely suppressed, a concept that, if proven, would fundamentally change our understanding of the disease. This suppression, effectively self-suppression, occurs in the context of the global inhibition of the cellular cap-dependent protein synthesis as a consequence of the neuronal integrated stress response (ISR) elicited by AβPP-derived intraneuronal Aβ (iAβ; hence self-suppression) upon reaching certain levels. Concurrently with the suppression of the AβPP proteolytic pathway, the neuronal ISR activates in human neurons, but not in mouse neurons, the powerful AD-driving pathway generating the C99 fragment of AβPP independently of AβPP. The present study describes molecular mechanisms potentially involved in these phenomena, propounds novel approaches to generate transgenic animal models of AD, advocates for the utilization of human neuronal cells-based models of the disease, makes verifiable predictions, suggests experiments designed to validate the proposed concept, and considers its potential research and therapeutic implications. Remarkably, it opens up the possibility that the conventional production of AβPP, BACE enzymes, and γ-secretase components is also suppressed under the neuronal ISR conditions in AD-affected neurons, resulting in the dyshomeostasis of AβPP. It follows that whereas conventional AD is triggered by AβPP-derived iAβ accumulated to the ISR-eliciting levels, the disease, in its both conventional and unconventional (triggered by the neuronal ISR-eliciting stressors distinct from iAβ) forms, is driven not (or not only) by iAβ produced in the AβPP-independent pathway, as we proposed previously, but mainly, possibly exclusively, by the C99 fragment generated independently of AβPP and not cleaved at the γ-site due to the neuronal ISR-caused deficiency of γ-secretase (apparently, the AD-driving “substance X” predicted in our previous study), a paradigm consistent with a dictum by George Perry that Aβ is “central but not causative” in AD. The proposed therapeutic strategies would not only deplete the driver of the disease and abrogate the AβPP-independent production of C99 but also reverse the neuronal ISR and ameliorate the AβPP dyshomeostasis, a potentially significant contributor to AD pathology. Full article
Show Figures

Figure 1

Figure 1
<p><b>AβPP-derived <span class="html-italic">i</span>Aβ cannot reach the AD pathology-causing range in human neurons</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: Levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: Cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: Levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Gradient-pink box</span>: Range of concentrations of <span class="html-italic">i</span>Aβ that cause and propel AD pathology. <span class="html-italic">Red box</span>: Apoptotic zone of <span class="html-italic">i</span>Aβ concentrations; within this zone the neurons have either undergone apoptosis or are dead. (<b>A</b>) The accumulation of AβPP-derived <span class="html-italic">i</span>Aβ is an exceedingly slow process; in the majority of humans, it neither reaches nor crosses the T1 threshold within their lifetime. (<b>B</b>) Provided the accumulation of <span class="html-italic">i</span>Aβ continues at the constant linear rate following the crossing of the T1 threshold, it cannot possibly reach the concentrations that cause and drive AD pathology and propel the progression of the disease.</p>
Full article ">Figure 2
<p><b>Conventional AD in the ACH2.0 perspective</b>. <span class="html-italic">i</span>Aβ: Intraneuronal Aβ. <span class="html-italic">eIF2α:</span> Eukaryotic translation initiation factor 2<span class="html-italic">α</span>. <span class="html-italic">PKR and HRI:</span> Kinases capable, when activated, of phosphorylating eIF2α. <span class="html-italic">TNFα</span>: Tumor necrosis factor α; potentially capable of activating PKR. <span class="html-italic">PACT</span>: PKR activator. <span class="html-italic">OMA1:</span> Mitochondrial distress-activated mitochondrial protease. <span class="html-italic">DELE1:</span> Substrate of OMA1; its cleavage leads to the activation of HRI. <span class="html-italic">nISR:</span> Neuronal integrated stress response; it is elicited by phosphorylation of eIF2α and enables the generation of components needed for the operation of the AβPP-independent pathway of <span class="html-italic">i</span>Aβ production. <span class="html-italic">AICD:</span> AβPP intracellular domain; results from the processing of C99. AβPP-derived <span class="html-italic">i</span>Aβ accrues physiologically in a life-long process. In most individuals, it does not reach the T1 threshold and no AD occurs. When the T1 threshold is crossed, PKR and/or HRI kinases are activated, eIF2α is phosphorylated, the ISR is elicited, the operation of the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway is initiated, and the disease commences. The bulk if not the entire Aβ output of the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway are retained as <span class="html-italic">i</span>Aβ. Its rapid accumulation both drives AD pathology and, by sustaining the activity of PKR and/or HRI, propagates the neuronal ISR and, consequently, perpetuates its own AβPP-independent generation. This cycle, denoted by the arched red and blue arrows, constitutes the driver of AD, the “AD Engine”.</p>
Full article ">Figure 3
<p><b>Dynamics of AβPP-derived <span class="html-italic">i</span>Aβ accumulation determines the occurrence and timing of conventional AD</b>. <span class="html-italic">i</span>Aβ: Intraneuronal Aβ. <span class="html-italic">Blue lines</span>: Levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: Cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: Levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: Apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. With the exception of the rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ, all parameters are considered constant. In these terms, the occurrence of AD is a function of and its timing is inversely proportional to the rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ. (<b>A</b>) The rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ is such that the T1 threshold is crossed and AD commences in the patient’s mid-sixties, statistically the age of onset of sporadic AD. In (<b>B</b>), the rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ decreases and the age of the T1 crossing and, consequently, of the onset of AD increases to the patient’s eighties. In (<b>C</b>), the rate of accumulation decreases further and the T1 threshold is crossed and the disease occurs in the patient’s nineties. (<b>D</b>) The rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ is such that the T1 threshold is not reached within the lifetime of the individual; neither is the T1 threshold crossed nor does conventional AD occur.</p>
Full article ">Figure 4
<p><b>The occurrence and timing of AD and the incidence of AACD are directly proportional to the extent of the T1 threshold</b>. <span class="html-italic">i</span>Aβ: Intraneuronal Aβ. <span class="html-italic">Blue lines</span>: Levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span><sup>0</sup></b> <span class="html-italic">threshold</span>: The level of AβPP-derived <span class="html-italic">i</span>Aβ above which the neuronal damage manifesting as AACD commences. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: Cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: Levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: Apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Gradient-pink boxes</span>: The range of AβPP-derived <span class="html-italic">i</span>Aβ concentrations between the T<sup>0</sup> and T1 thresholds (provided T<sup>0</sup> &lt; T1), a zone of the occurrence of AACD. All parameters (including the rate of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ and the extent of the T<sup>0</sup> threshold), with the exception of the extent of the T1 threshold, are considered constant. (<b>A</b>) The extent of the T1 threshold is smaller than that of the T<sup>0</sup> threshold, such that the T1 crossing occurs and AD commences at the age of sixty; no AACD occurs. (<b>B</b>) The extent of the T1 threshold increases and is greater than that of the T<sup>0</sup> threshold. AACD commences when the T<sup>0</sup> is reached and morphs into AD with the T1 crossing at about 75 years. (<b>C</b>) The T1 threshold is elevated further. AACD commences at the same time as in panel B but the T1 crossing and the commencement of AD occur later, at about ninety years. (<b>D</b>) The extent of the T1 threshold is elevated further; neither is it crossed nor does AD occur. AACD commences at precisely the same time as in (<b>B</b>,<b>C</b>) and persists for the remaining lifespan.</p>
Full article ">Figure 5
<p><b>Principal stages of the chimeric pathway of mammalian RNA-dependent mRNA amplification</b>. <span class="html-italic">RdRp</span>: RNA-dependent RNA polymerase. <span class="html-italic">Single line</span>: Antisense RNA. <span class="html-italic">Boxed line</span>: Sense RNA. <span class="html-italic">AUG</span>: Translation initiation codon. <span class="html-italic">TCE</span>: 3′ Terminal Complementary Element of the antisense RNA. <span class="html-italic">ICE</span>: Internal Complementary Element of the antisense RNA. <span class="html-italic">Yellow circle</span>: A complex containing helicase, nucleotide-modifying, and RNA-cleaving activities. <span class="html-italic">Blue lines</span>: Separated single-stranded RNA molecules or portions thereof. <span class="html-italic">Red arrows</span>: The site of cleavage of the chimeric RNA intermediate. <span class="html-italic">Top Panel</span>: A progenitor of the RNA-dependent mammalian mRNA amplification—conventionally genome-transcribed mRNA molecule. <span class="html-italic">Middle Panel</span>: Principal stages of the chimeric pathway of the mRNA amplification process. Stage <b>1</b>: Antisense RNA is transcribed by RdRp from the progenitor mRNA. Stage <b>2</b>: Sense and antisense RNA molecules are separated by the helicase activity. Helicase mounts the 3′ poly(A) segment and moves along the sense orientation molecule modifying, on average, every fifth nucleotide. Stage <b>3</b>: TCE/ICE interaction-guided folding of the antisense RNA into self-priming configuration. Stage <b>4</b>: RdRp extends the 3′ end of self-primed antisense RNA; the result is a hairpin-like chimeric molecule containing sense and antisense RNA components and referred to as the chimeric RNA intermediate. Stage <b>5</b>: The double-stranded portion of the chimeric RNA intermediate is separated by helicase; nucleotide modifications introduced during the separation prevent re-annealing of the RNA strands. Stage <b>6</b>: The helicase complex cleaves the chimeric RNA intermediate upon reaching its single-stranded portion. Stage <b>7</b>: Chimeric RNA and antisense RNA end products. The antisense RNA is 3′-truncated, and its cleaved-off segment becomes the 5′ terminus of the chimeric RNA end product. Note: In this scenario, the ICE is situated within a portion of the antisense RNA complementary to a segment of the 5′UTR of the progenitor mRNA. Therefore, the chimeric RNA end product retains the entire coding region of the progenitor. <span class="html-italic">Bottom Panel</span>: The ICE is located within a portion of the antisense RNA complementary to a segment of the coding region of the progenitor mRNA. The chimeric RNA end product is 5′-truncated within the coding region of the progenitor. The potential outcome of its translation is defined by the position of the first functional translation-initiating codon. If it is located within the retained portion of the coding region and is in frame, the outcome would be the CTF of the progenitor mRNA-encoded protein. Stages <b>3′</b>–<b>7′</b> of the bottom panel correspond to stages <b>3</b>–<b>7</b> of the middle panel.</p>
Full article ">Figure 6
<p><b>Human AβPP mRNA is an eligible template of asymmetric amplification; the resulting mRNA encodes the C100 fragment of AβPP</b>. <span class="html-italic">Lowercase letters</span>: nucleotide sequences of the portions of the human antisense AβPP RNA capable of forming the self-priming structure. <span class="html-italic">Uppercase letters</span>: nucleotide sequence of a segment of human AβPP mRNA generated by the extension of self-primed human antisense AβPP RNA. <span class="html-italic">Highlighted in yellow</span>: the TCE (3′-terminal segment) and the ICE (internal segment) elements of the antisense RNA. <span class="html-italic">2011–2013</span>: Positions of nucleotides counted from the 3′ terminus of the antisense RNA and forming the “UAC” (<span class="html-italic">highlighted in blue</span>) complementary to the AUG codon (<span class="html-italic">highlighted in green</span>) encoding Met671 of human AβPP. (<b>a</b>–<b>c</b>) parallel stages 3′ to 7′ of <a href="#genes-16-00046-f005" class="html-fig">Figure 5</a>. (<b>a</b>) TCE/ICE interaction-guided folding of human antisense AβPP RNA into self-priming configuration. (<b>b</b>) RdRp-mediated extension of the 3′ terminus of self-primed human antisense AβPP RNA; the resulting segment of human AβPP mRNA is <span class="html-italic">highlighted in gray</span>. When, upon separating the RNA strands, helicase reaches the single-stranded portion of the chimeric RNA intermediate, it cleaves either at its 3′ end (<span class="html-italic">red arrow</span>) or at a mismatch within the TCE/ICE structure. (<b>c</b>) The chimeric RNA end product (<span class="html-italic">highlighted in gray</span>) consists of AβPP mRNA 5′-truncated within its coding region and containing a portion of the antisense RNA appended to its 5′ end. Its first, 5′-most translation initiation codon is the AUG encoding methionine 671 of human AβPP.</p>
Full article ">Figure 7
<p><b>Mouse AβPP mRNA is ineligible for RNA-dependent mRNA amplification</b>. (<b>Top</b>) Human antisense AβPP RNA folded into self-priming structure. (<b>Bottom</b>) The relationship between the portions of the mouse antisense AβPP RNA in positions analogous to those of the human TCE and ICE elements. <span class="html-italic">Asterisks</span> denote the nucleotide position corresponding to the transcription start site of either human or mouse AβPP mRNA; in both cases, it is the nucleotide (-)149 upstream from the AUG translation initiation codon. The relationship between segments on mouse antisense AβPP RNA corresponding to the TCE and ICE elements of its human counterpart is, in its strength, no better than random. Moreover, the occurrence of the 3′ overhang in the “folded” mouse antisense AβPP RNA is incompatible with the priming function.</p>
Full article ">Figure 8
<p><b>Projected dynamics of accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in transgenic mouse models overexpressing human AβPP. <span class="html-italic">i</span>Aβ</b>: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: Cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Gradient-pink box</span>: Range of concentrations of <span class="html-italic">i</span>Aβ that cause and propel AD pathology. <span class="html-italic">Red box</span>: Apoptotic zone; within this zone the neurons have either undergone apoptosis or are dead. (<b>A</b>) Accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in a wild mouse. The rate of accumulation is exceedingly slow. AβPP-derived <span class="html-italic">i</span>Aβ neither reaches nor crosses the T1 threshold within the lifetime of the mouse. (<b>B</b>) Projected accumulation of exogenous human AβPP-derived <span class="html-italic">i</span>Aβ in transgenic mouse models. For reasons given in the main text, the rate of accumulation of exogenous AβPP-derived <span class="html-italic">i</span>Aβ should be on par with that of <span class="html-italic">i</span>Aβ produced independently of AβPP in AD patients. It should reach the AD pathology-causing range, cross the T2 threshold, and elicit full symptomatic spectrum of AD. This, apparently, does not happen.</p>
Full article ">Figure 9
<p><b>Production of Aβ in the AβPP proteolytic pathway is either discontinued or severely suppressed in both humans and mouse under the neuronal ISR following the T1 crossing</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of AβPP-derived <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">Red lines</span>: Levels of <span class="html-italic">i</span>Aβ generated independently of AβPP in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: Cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Gradient-pink boxes</span>: Range of concentrations of <span class="html-italic">i</span>Aβ that cause and propel AD pathology. <span class="html-italic">Red boxes</span>: Apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. (<b>A</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ in the AD patient. AβPP-derived <span class="html-italic">i</span>Aβ accumulates physiologically as described in the main text and crosses the T1 threshold. This triggers the activation of PKR and/or HRI kinases, phosphorylation of eIF2α at its Ser51 residue, and the elicitation of the neuronal ISR. Under the ISR conditions, global cellular translation is reprogrammed and protein production is severely suppressed. It is presumed that the production of Aβ in the AβPP proteolytic pathway is no exception. Consequently, following the crossing of the T1 threshold by the AβPP-derived <span class="html-italic">i</span>Aβ, the rate of its accumulation substantially declines. Concurrently, the neuronal ISR activates and sustains the generation of <span class="html-italic">i</span>Aβ in the AβPP-independent pathway. It rapidly accumulates, reaches the AD pathology-causing range, and AD symptoms manifest. When it crosses the T2 threshold, the disease enters its end stage. (<b>B</b>) Dynamics of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in transgenic mouse models. Both endogenous and exogenous AβPP-derived <span class="html-italic">i</span>Aβ accumulate physiologically via the cellular uptake of a fraction of secreted extracellular Aβ and the retention of <span class="html-italic">i</span>Aβ resulting from a fraction of C99 processed on the intraneuronal membranes. When the T1 threshold is crossed, the neuronal ISR is elicited as described above. The suppression of the production of Aβ in the AβPP proteolytic pathway, occurring in the context of the global suppression of the cellular protein synthesis, results in a substantial decline in the rate of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ. Since the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway is inoperative in both wild mice and mouse models, the AD pathology-causing range of <span class="html-italic">i</span>Aβ concentrations would not be reached and AD symptoms would not occur.</p>
Full article ">Figure 10
<p><b>Inhibition of the translation of AβPP as the potential origin of the neuronal ISR-mediated suppression of the production of AβPP-derived <span class="html-italic">i</span>Aβ</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of AβPP-derived <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">Red lines</span>: levels of <span class="html-italic">i</span>Aβ generated independently of AβPP in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Gradient-pink boxes</span>: range of concentrations of <span class="html-italic">i</span>Aβ that cause and propel AD pathology. <span class="html-italic">Red boxes</span>: apoptotic zone; within this zone the neurons have either undergone apoptosis or are dead. (<b>A</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ in AD patients. AβPP-derived <span class="html-italic">i</span>Aβ accumulates physiologically (as described in the main text) and crosses the T1 threshold. This triggers the activation of PKR and/or HRI kinases, phosphorylation of eIF2α at its Ser51 residue, and the elicitation of the neuronal ISR. Under the ISR conditions, global cellular translation is reprogrammed and protein production is severely suppressed. It is presumed that the production of AβPP and, consequently, of Aβ in the AβPP proteolytic pathway is no exception. Consequently, following the crossing of the T1 threshold by the AβPP-derived <span class="html-italic">i</span>Aβ, the rate of its accumulation is reversed. Concurrently, the neuronal ISR activates and sustains the generation of <span class="html-italic">i</span>Aβ in the AβPP-independent pathway. Thus, although the levels of AβPP-derived <span class="html-italic">i</span>Aβ decrease below the T1 threshold, the ISR state, maintained now by <span class="html-italic">i</span>Aβ produced independently of AβPP, remains in effect. (<b>B</b>) Dynamics of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in transgenic mouse models. Both endogenous and exogenous AβPP-derived <span class="html-italic">i</span>Aβ accumulate physiologically via the cellular uptake of a fraction of secreted extracellular Aβ and the retention of <span class="html-italic">i</span>Aβ resulting from a fraction of C99 processed on the intraneuronal membranes. When the T1 threshold is crossed, the neuronal ISR is elicited as described above. The suppression of the production of AβPP and, consequently, of Aβ in the AβPP proteolytic pathway, in the context of the global suppression of the cellular protein synthesis, results in reversing the rate of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ. Since the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway is inoperative in both wild mice and mouse models, the reverse crossing of the T1 threshold restores normal (i.e., non-ISR) conditions. The production of AβPP and, consequently, the influx of AβPP-derived <span class="html-italic">i</span>Aβ resumes. Eventually, the latter crosses the T1 threshold, triggers the re-elicitation of the neuronal ISR, and the oscillating cycle repeats.</p>
Full article ">Figure 11
<p><b>The neuronal ISR-caused deficiency of γ-secretase as the basis of the suppression of the production of AβPP-derived <span class="html-italic">i</span>Aβ</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">C99</span>: The fragment of AβPP produced either by the proteolysis of AβPP or independently of AβPP. <span class="html-italic">Blue lines</span>: levels of AβPP-derived <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">Red lines</span>: levels of the C99 fragment of AβPP in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ or of C99 that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ or C99 that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Gradient-pink boxes</span>: range of concentrations of C99 or <span class="html-italic">i</span>Aβ that cause and propel AD pathology. <span class="html-italic">Red boxes</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. (<b>A</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ and C99 in the AD patient. AβPP-derived <span class="html-italic">i</span>Aβ accumulates physiologically, as described in the main text, and crosses the T1 threshold. This triggers the activation of PKR and/or HRI kinases, phosphorylation of eIF2α at its Ser51 residue, and the elicitation of the neuronal ISR. Under the ISR conditions, global cellular translation is reprogrammed and protein production is severely suppressed. It is presumed that the production of the components of the AβPP proteolytic pathway in general and that of γ-secretase in particular is no exception. Consequently, following the crossing of the T1 threshold by the AβPP-derived <span class="html-italic">i</span>Aβ, the rate of its accumulation declines (as shown) or is reversed. Concurrently, the neuronal ISR activates and sustains the operation of the AβPP-independent pathway, which generates, due to the deficiency of γ-secretase, C99. The latter rapidly accumulates, crosses the T1 threshold, reaches the AD pathology-causing range, and eventually crosses the T2 threshold. It both drives the disease and propagates the neuronal ISR state, thus perpetuating its own production. (<b>B</b>) The dynamics of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in transgenic mouse models. Both endogenous and exogenous AβPP-derived <span class="html-italic">i</span>Aβ accumulate physiologically via the cellular uptake of a fraction of secreted extracellular Aβ and the retention of <span class="html-italic">i</span>Aβ resulting from a fraction of C99 processed on the intraneuronal membranes. When the T1 threshold is crossed, the neuronal ISR is elicited as described above. The suppression of the production of Aβ in the AβPP proteolytic pathway in the context of the global inhibition of cellular protein synthesis results in a substantial decline in the rate of the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ. Since the AβPP-independent C99 generation pathway is inoperative in both wild mice and mouse models, the AD pathology-causing range would be reached neither by <span class="html-italic">i</span>Aβ nor by C99, and AD symptoms would not occur. If the rate of AβPP-derived <span class="html-italic">i</span>Aβ were reversed under the neuronal ISR conditions, the oscillating pattern of <span class="html-italic">i</span>Aβ levels shown in <a href="#genes-16-00046-f010" class="html-fig">Figure 10</a>B would result.</p>
Full article ">Figure 12
<p>Long-term suppression of the integrated stress response in the prevention of conventional Alzheimer’s disease. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span><sup>0</sup></b> <span class="html-italic">threshold</span>: the level of AβPP-derived <span class="html-italic">i</span>Aβ above which the neuronal damage manifesting as AACD commences. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the inhibition of the ISR. <span class="html-italic">Gradient-pink boxes</span>: the range of AβPP-derived <span class="html-italic">i</span>Aβ concentrations between the T<sup>0</sup> and T1 thresholds (provided T<sup>0</sup> &lt; T1), a zone of the occurrence of AACD; note that AACD can persist over the T1 threshold if the commencement of AD is prevented. (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons; they are below both T<sup>0</sup> and T1. (<b>B</b>) Evolution of the initial state in the untreated person. AβPP-derived <span class="html-italic">i</span>Aβ reaches the T<sup>0</sup> threshold and AACD commences. With the crossing of the T1 threshold the neuronal ISR is elicited, the AβPP-independent C100/C99 generation pathway activated, and AD commences. At this point AACD morphs into AD. <span class="html-italic">i</span>Aβ produced independently of AβPP accumulates, reaches the T2 threshold and the disease enters the end stage. (<b>C</b>) Evolution of the initial state in the presence of ISR inhibitors. With the crossing of the T<sup>0</sup> threshold by AβPP-derived <span class="html-italic">i</span>Aβ AACD commences. When the T1 threshold is crossed the neuronal ISR cannot be elicited and the AβPP-independent C100/C99 generation pathway remains inactive. No AD occurs, but AACD persists for the duration of the treatment.</p>
Full article ">Figure 13
<p>Long-term suppression of the integrated stress response in the treatment of conventional Alzheimer’s disease. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the inhibition of the ISR. (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons. AβPP-derived <span class="html-italic">i</span>Aβ has already crossed the T1 threshold in all affected neurons of the AD patient. Consequently, the neuronal integrated stress response is elicited, the AβPP-independent C100/C99 generation pathway is activated, and <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates. The apoptosis-triggering T2 threshold is crossed in a fraction of the neurons, but the bulk of the affected neurons remain sub-T2. (<b>B</b>) Evolution of the initial state in the untreated individual. The AβPP-independent C100/C99 generation pathway remains operational, the accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP continues uninterrupted, and more neurons cross the T2 threshold, leading to the end stage of the disease. <span class="html-italic">i</span>Aβ produced independently of AβPP accumulates, reaches the T2 threshold, and the disease enters the end stage. (<b>C</b>) Evolution of the initial state in the presence of ISR inhibitors. With the supply of essential components interrupted, the operation of the AβPP-independent C100/C99 generation pathway ceases. However, due to the influx of AβPP-derived <span class="html-italic">i</span>Aβ, the levels of total <span class="html-italic">i</span>Aβ continue to increase, although at the slow, pre-T1 crossing rate. Levels of <span class="html-italic">i</span>Aβ produced independently of AβPP continue to cross the T2 threshold, and the progression of the disease would persist albeit at a decreased rate for the duration of the treatment.</p>
Full article ">Figure 14
<p><b>Transient administration of the ISR inhibitors is feasible but inefficient in both prevention and treatment of conventional AD. <span class="html-italic">i</span>Aβ: intraneuronal Aβ</b>. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the inhibition of the ISR. (<b>A</b>) Transient administration of ISR inhibitors prior to the crossing of the T1 threshold and commencement of AD. At this time, the neuronal ISR has not yet been elicited, and the presence of its inhibitors would have no beneficial effect; the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ would continue uninterrupted. (<b>B</b>) Transient administration of ISR inhibitors in the treatment of conventional AD. The neuronal ISR has been elicited and the AβPP-independent C100/C99 production pathway rendered operational in all affected neurons. Inhibition of the neuronal ISR interrupts the production of components required for the activity of this pathway and its operation ceases for the duration of the treatment. However, <span class="html-italic">i</span>Aβ continues to accumulate at the pre-T1 crossing rate due to the influx of AβPP-derived <span class="html-italic">i</span>Aβ. When the transient treatment is concluded, <span class="html-italic">i</span>Aβ levels in all affected neurons remain at the over-T1 levels. Consequently, the ISR is re-elicited, the operation of the AβPP-independent C100/C99 generation pathway is restored, and the accumulation of <span class="html-italic">i</span>Aβ and progression of AD resumes at the pre-treatment rate; the benefits of the treatment are thus limited to the duration of its administration.</p>
Full article ">Figure 15
<p><b>The depletion of <span class="html-italic">i</span>Aβ would be very effective in the prevention of conventional AD and AACD and in the treatment of AACD. <span class="html-italic">i</span>Aβ: intraneuronal Aβ</b>. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span><sup>0</sup></b> <span class="html-italic">threshold</span>: the level of AβPP-derived <span class="html-italic">i</span>Aβ above which the neuronal damage manifesting as AACD commences. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Orange boxes</span>: Duration of the depletion of AβPP-derived <span class="html-italic">i</span>Aβ. <span class="html-italic">Gradient-pink boxes</span>: The range of AβPP-derived <span class="html-italic">i</span>Aβ concentrations between the T<sup>0</sup> and T1 thresholds (provided T<sup>0</sup> &lt; T1), a zone of the occurrence of AACD. (<b>A</b>,<b>B</b>) The extent of the T<sup>0</sup> is greater than that of the T1 threshold; no AACD can occur. (<b>A</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ and progression of the disease in the untreated AD patient. AβPP-derived <span class="html-italic">i</span>Aβ reaches the T1 threshold and triggers the elicitation of the neuronal ISR and, consequently, the activation of the AβPP-independent C100/C99 generation pathway. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates and crosses the apoptosis-inducing T2 threshold. (<b>B</b>) The transient <span class="html-italic">i</span>Aβ depletion treatment is administered prior to the crossing of the T1 threshold and commencement of AD. AβPP-derived <span class="html-italic">i</span>Aβ is substantially depleted and its de novo accumulation resumes from a low baseline and proceeds at the pre-treatment rate. It does not reach the T1 threshold and AD does not occur within the lifetime of the treated individual. (<b>C</b>,<b>D</b>) The extent of the T1 is greater than that of the T<sup>0</sup> threshold; AD is preceded by AACD. (<b>C</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ and progression of the disease in the untreated AACD/AD patient. AβPP-derived <span class="html-italic">i</span>Aβ reaches the T<sup>0</sup> threshold and AACD commences. AβPP-derived <span class="html-italic">i</span>Aβ continues to accumulate and when it crosses the T1 threshold, the neuronal ISR is elicited, AβPP-independent C100/C99 generation pathway activated, AD commences, and AACD morphs into AD. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates, crosses into the apoptotic zone, and the disease enters the end stage. (<b>D</b>) The transient <span class="html-italic">i</span>Aβ depletion treatment is administered when the T<sup>0</sup> threshold has already been crossed and AACD commenced but prior to the T1 crossing. AβPP-derived <span class="html-italic">i</span>Aβ is substantially depleted. The progression of AACD ceases and the condition is cured. AβPP-derived <span class="html-italic">i</span>Aβ resumes its de novo accumulation from a low baseline. It does not reach the T<sup>0</sup> threshold within the remaining lifetime of the treated individual; neither AACD recurs nor AD occurs. Note that if the <span class="html-italic">i</span>Aβ depletion treatment were administered prior to the crossing of the T<sup>0</sup> threshold, both AACD and conventional AD would be prevented.</p>
Full article ">Figure 16
<p><b>The depletion of <span class="html-italic">i</span>Aβ would be inefficient in the treatment of AD. <span class="html-italic">i</span>Aβ: intraneuronal Aβ</b>. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <b><span class="html-italic">T</span>1</b> <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <b><span class="html-italic">T</span>2</b> <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Orange box</span>: duration of the depletion of AβPP-derived <span class="html-italic">i</span>Aβ. (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons. AβPP-derived <span class="html-italic">i</span>Aβ has already crossed the T1 threshold in all affected neurons of the AD patient. The neuronal integrated stress response has been elicited, the AβPP-independent C100/C99 generation pathway has been activated, and <span class="html-italic">i</span>Aβ produced independently of AβPP has rapidly accumulated. The apoptosis-triggering T2 threshold has been crossed in a fraction of the neurons, but the bulk of the affected neurons remain sub-T2. (<b>B</b>) Evolution of the initial state in the untreated AD patient. The AβPP-independent C100/C99 generation pathway remains operational, the accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP continues uninterrupted, and more neurons cross the T2 threshold, thus leading to the end stage of the disease. (<b>C</b>) Evolution of the initial state in the presence of BACE1 and/or BACE2 activators. The rate of the influx of <span class="html-italic">i</span>Aβ produced in the AβPP-independent pathway outbalances that of the removal of <span class="html-italic">i</span>Aβ by intra-<span class="html-italic">i</span>Aβ cleaving activities of activated BACE1 and/or BACE2. Its accumulation as well as the progression of the disease persists, although at a decreased rate, for the duration of the treatment.</p>
Full article ">Figure 17
<p><b>Composite therapeutic strategy for conventional symptomatic AD: concurrent transient administration of ISR inhibitors and BACE1 and/or BACE2 activators</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the administration of ISR inhibitors. <span class="html-italic">Orange box</span>: duration of the depletion of AβPP-derived <span class="html-italic">i</span>Aβ: (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons. AβPP-derived <span class="html-italic">i</span>Aβ has already crossed the T1 threshold in all affected neurons of the AD patient. The neuronal integrated stress response is elicited, the AβPP-independent C100/C99 generation pathway is activated, and <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates. The apoptosis-triggering T2 threshold crosses in a fraction of the neurons, but the bulk of the affected neurons remain sub-T2. (<b>B</b>) Evolution of the initial state in the untreated individual. The AβPP-independent C100/C99 generation pathway remains operational, the accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP continues uninterrupted, and more neurons cross the T2 threshold, leading to the end stage of the disease. <span class="html-italic">i</span>Aβ produced independently of AβPP accumulates, reaches the T2 threshold, and the disease enters the end stage. (<b>C</b>) Evolution of the initial state in the treated AD patient. The transient ISR inhibition and BACE activation are implemented concurrently. The former enables the production of BACE1 and BACE2 and thus ensures their availability, and disables the AβPP-independent C100/C99 generation pathway, thus abolishing the influx of <span class="html-italic">i</span>Aβ produced independently of AβPP and ensuring its efficient depletion by the latter. Following the treatment, the neuronal ISR state is reversed to normal, the AβPP-independent C100/C99 generation pathway is left inoperative, and AβPP-derived <span class="html-italic">i</span>Aβ resumes its accumulation de novo from a low baseline. It does not reach the T1 threshold, and AD does not recur within the remaining lifetime of the treated patient.</p>
Full article ">Figure 18
<p><b>Conventional and unconventional AD differ only in the manner of the elicitation of the neuronal ISR.</b> <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">AβPP</span>: Aβ protein precursor. <span class="html-italic">eIF2α:</span> eukaryotic translation initiation factor 2<span class="html-italic">α</span>. <span class="html-italic">PKR</span> and <span class="html-italic">HRI:</span> kinases capable of phosphorylating eIF2α when activated. <span class="html-italic">TNFα</span>: tumor necrosis factor α, potentially capable of activating PKR. <span class="html-italic">PACT</span>: PKR activator. <span class="html-italic">OMA1:</span> mitochondrial distress-activated mitochondrial protease. <span class="html-italic">DELE1:</span> substrate of OMA1; its cleavage leads to the activation of HRI. <span class="html-italic">ISR:</span> neuronal integrated stress response; it is elicited by phosphorylation of eIF2α and enables the production of components essential for the operation of the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway. <span class="html-italic">AICD:</span> AβPP intracellular domain; it results from the processing of C99. <span class="html-italic">AD Engine</span> (denoted by the arched blue and red arrows): repeated feedback cycles of the propagation of the ISR state by <span class="html-italic">i</span>Aβ generated independently of AβPP and the resulting stimulation of its own production. In conventional AD, the elicitation of the neuronal ISR is triggered via activation of PKR and/or HRI kinases and phosphorylation of eIF2α, by AβPP-derived <span class="html-italic">i</span>Aβ accumulated over the T1 threshold (designated “conventional stressor”). In unconventional AD, the elicitation of the neuronal ISR is triggered via activation of one or more eIF2α kinases, by stressors other than AβPP-derived <span class="html-italic">i</span>Aβ (designated “unconventional” stressors). In both forms of AD, the neuronal ISR enables, via the supply of essential components, the operation of the AβPP-independentC100/C99 generation pathway. The resulting <span class="html-italic">i</span>Aβ, produced independently of AβPP, drives AD pathology, propagates the neuronal ISR, and perpetuates its own production. In contrast to conventional AD, where the AβPP-independent C100/C99 generation pathway is self-sustainable from the instance of its activation (because <span class="html-italic">i</span>Aβ levels are already above the T1 threshold), in unconventional AD, the AβPP-independent C100/C99 generation pathway is activated at the levels of AβPP-derived <span class="html-italic">i</span>Aβ below the T1 threshold and becomes self-sustainable only when its <span class="html-italic">i</span>Aβ product crosses the latter. In this scenario, the accumulation of <span class="html-italic">i</span>Aβ produced in the AβPP-independent pathway is accompanied by a corresponding accumulation of AICD. AICD was demonstrated to be capable of interfering with various AD components but its potential contribution to the disease remains to be elucidated [<a href="#B237-genes-16-00046" class="html-bibr">237</a>,<a href="#B238-genes-16-00046" class="html-bibr">238</a>,<a href="#B239-genes-16-00046" class="html-bibr">239</a>,<a href="#B240-genes-16-00046" class="html-bibr">240</a>,<a href="#B241-genes-16-00046" class="html-bibr">241</a>,<a href="#B242-genes-16-00046" class="html-bibr">242</a>,<a href="#B243-genes-16-00046" class="html-bibr">243</a>,<a href="#B244-genes-16-00046" class="html-bibr">244</a>,<a href="#B245-genes-16-00046" class="html-bibr">245</a>,<a href="#B246-genes-16-00046" class="html-bibr">246</a>,<a href="#B247-genes-16-00046" class="html-bibr">247</a>,<a href="#B248-genes-16-00046" class="html-bibr">248</a>,<a href="#B249-genes-16-00046" class="html-bibr">249</a>,<a href="#B250-genes-16-00046" class="html-bibr">250</a>,<a href="#B251-genes-16-00046" class="html-bibr">251</a>,<a href="#B252-genes-16-00046" class="html-bibr">252</a>].</p>
Full article ">Figure 19
<p><b>Dynamics of iAb in unconventional AD: Effect of the long-duration occurrence of unconventional stressors capable of the elicitation of the neuronal ISR</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Pink box</span>: duration of the presence of unconventional stressors capable of triggering the elicitation of the neuronal ISR: (<b>A</b>) Dynamics of the accumulation of <span class="html-italic">i</span>Aβ in a healthy person. <span class="html-italic">i</span>Aβ is produced solely by the proteolysis of AβPP; its levels do not reach the T1 threshold and the unconventional stressors do not occur within the lifetime of the individual. (<b>B</b>) Unconventional stressors occur when levels of AβPP-derived <span class="html-italic">i</span>Aβ are below the T1 threshold and persist for the remaining lifetime. The neuronal ISR is unconventionally elicited, and the AβPP-independent C100/C99 production pathway is activated. When <span class="html-italic">i</span>Aβ produced in this pathway crosses the T1 threshold, the pathway becomes self-sustainable, and AD commences and progresses to its end stage.</p>
Full article ">Figure 20
<p><b>Dynamics of <span class="html-italic">i</span>Aβ in unconventional AD: Effect of the transient occurrence of unconventional stressors capable of the elicitation of the neuronal ISR</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">Broken lines</span>: projected accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in the absence of unconventional stressors. <span class="html-italic">T</span><sup>0</sup> <span class="html-italic">threshold</span>: the level of AβPP-derived <span class="html-italic">i</span>Aβ above which the neuronal damage manifesting as AACD commences. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Pink box</span>: duration of the presence of unconventional stressors capable of triggering the elicitation of the neuronal ISR. <span class="html-italic">Gradient-pink boxes</span>: the range of AβPP-derived <span class="html-italic">i</span>Aβ concentrations between the T<sup>0</sup> and T1 thresholds (provided T<sup>0</sup> &lt; T1), a zone of the occurrence of AACD. (<b>A</b>) The unconventionally activated AβPP-independent C100/C99 generation pathway is operational for only a short duration, insufficient for the T1 crossing by <span class="html-italic">i</span>Aβ. When unconventional stressors are withdrawn, <span class="html-italic">i</span>Aβ, produced solely by AβPP proteolysis, continues to accumulate at a slow rate. It does not reach the T1 threshold, and no AD occurs within the lifespan of the individual. It does, however, cross the T<sup>0</sup> threshold; AACD commences and persists for the remaining lifetime of the individual. (<b>B</b>) The unconventionally activated AβPP-independent C100/C99 production pathway operates sufficiently long for its <span class="html-italic">i</span>Aβ product to cross both the T<sup>0</sup> and T1 thresholds. Upon the T<sup>0</sup> crossing AACD commences, and when the T1 is crossed, the AβPP-independent C100/C99 production pathway is rendered self-sustainable, and AACD morphs into AD. (<b>C</b>) The unconventionally activated AβPP-independent C100/C99 production pathway operates long enough for its <span class="html-italic">i</span>Aβ product to cross the T<sup>0</sup> and T1 thresholds in only a portion of the affected neurons. Upon the T<sup>0</sup> crossing, AACD commences and persists until the T1 crossing. When unconventional stressors are withdrawn, the operation of the AβPP-independent C100/C99 generation pathway in neurons that cross the T1 threshold continues uninterrupted and self-sustainably. In the rest of the affected neurons, the operation of the unconventionally activated AβPP-independent C100/C99 production pathway ceases. The accumulation of AβPP-derived <span class="html-italic">i</span>Aβ in these neurons continues at a slow rate; when it crosses the T1 threshold, the AβPP-independent C100/C99 generation pathway is conventionally activated, and the progression of AD pathology commences.</p>
Full article ">Figure 21
<p><b>Dynamics of iAβ in unconventional AD: Effect of the recurrent transient occurrence of unconventional stressors capable of the elicitation of the neuronal ISR</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Pink box</span>: duration of the presence of unconventional stressors capable of triggering the elicitation of the neuronal ISR: (<b>A</b>) Transient operation of the unconventionally activated AβPP-independent C100/C99 generation pathway accelerates the occurrence of conventional AD. The duration of the operation of the unconventionally activated AβPP-independent C100/C99 production pathway is such that the resulting accumulation of its <span class="html-italic">i</span>Aβ product is substantial, but its levels are still below the T1 threshold. When unconventional stressors are withdrawn, the operation of the AβPP-independent pathway ceases, but AβPP-derived <span class="html-italic">i</span>Aβ continues to accumulate at a slow rate but from a significantly elevated baseline. When it crosses the T1 threshold, the neuronal ISR is conventionally re-elicited, the AβPP-independent C100/C99 generation pathway is reactivated, and AD commences. (<b>B</b>) Effect of the recurrent transient occurrence of unconventional stressors capable of the elicitation of the neuronal ISR. The unconventional elicitation of the neuronal integrated stress response and the operation of the unconventionally activated AβPP-independent C100/C99 production pathway occurs for a short duration but recurrently. <span class="html-italic">i</span>Aβ accumulates in a stepwise fashion with incremental rounds of fast accumulation, resulting from the transient operation of unconventionally activated AβPP-independent C100/C99 production pathway, interspersed by the accumulation of AβPP-derived <span class="html-italic">i</span>Aβ occurring at a slow rate but from repeatedly elevated baselines. This accelerates the T1 crossing, and when it occurs, the self-sustainable AβPP-independent C100/C99 generation pathway is activated, and AD commences.</p>
Full article ">Figure 22
<p><b>Composite therapeutic strategy for the prevention of unconventional AD: concurrent transient administration of ISR inhibitors and BACE1 and/or BACE2 activators</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the administration of ISR inhibitors. <span class="html-italic">Orange box</span>: duration of the depletion of AβPP-derived <span class="html-italic">i</span>Aβ: (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons. The neuronal integrated stress response is unconventionally elicited and, consequently, the AβPP-independent C100/C99 production pathway is activated. Levels of <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly increase but have not yet crossed the T1 threshold. (<b>B</b>) Evolution of the initial state in the untreated individual. The AβPP-independent C100/C99 generation pathway remains operational, and the accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP continues uninterrupted; more neurons cross the T2 threshold, thus leading to the end stage of the disease. (<b>C</b>) Evolution of the initial state in the treated AD patient. The transient ISR inhibition and BACE activation are implemented concurrently. The former enables the production of BACE1 and BACE2 and thus ensures their availability, and disables the AβPP-independent C100/C99 generation pathway, thus abolishing the influx of <span class="html-italic">i</span>Aβ produced independently of AβPP and ensuring its efficient depletion by the latter. Following the treatment, the accumulation of <span class="html-italic">i</span>Aβ resumes from a low baseline. At this time, however, the neuronal integrated stress response is unconventionally elicited, due to the persistence of unconventional stressors, and the AβPP-independent C100/C99 generation pathway is unconventionally activated. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates and crosses the T1 threshold. The disease commences and proceeds unimpeded. The treatment delays the occurrence of AD but only transiently.</p>
Full article ">Figure 23
<p><b>Composite therapeutic strategy for the treatment of unconventional AD: concurrent transient administration of ISR inhibitors and BACE1 and/or BACE2 activators</b>. <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Green box</span>: duration of the administration of ISR inhibitors. <span class="html-italic">Orange box</span>: duration of the depletion of AβPP-derived <span class="html-italic">i</span>Aβ: (<b>A</b>) Initial state of AβPP-derived <span class="html-italic">i</span>Aβ levels in individual neurons. AβPP-derived <span class="html-italic">i</span>Aβ has already crossed the T1 threshold in all affected neurons of the AD patient. The neuronal integrated stress response is elicited, the AβPP-independent C100/C99 generation pathway is activated, and <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates. The apoptosis-triggering T2 threshold is crossed in a fraction of the neurons, but the bulk of the affected neurons remain sub-T2. (<b>B</b>) Evolution of the initial state in the untreated individual. The AβPP-independent C100/C99 generation pathway remains operational, and the accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP continues uninterrupted; more neurons cross the T2 threshold, thus leading to the end stage of the disease. <span class="html-italic">i</span>Aβ produced independently of AβPP accumulates and reaches the T2 threshold, and the disease enters the end stage. (<b>C</b>) Evolution of the initial state in the treated AD patient. The transient ISR inhibition and BACE activation are implemented concurrently. The former enables the production of BACE1 and BACE2 and thus ensures their availability, and disables the AβPP-independent C100/C99 generation pathway, thus abolishing the influx of <span class="html-italic">i</span>Aβ produced independently of AβPP and ensuring its efficient depletion by the latter. Following the treatment, the progression of AD ceases, and the accumulation of <span class="html-italic">i</span>Aβ resumes from a low baseline. At this time, however, the neuronal integrated stress response is unconventionally elicited, due to the persistence of unconventional stressors, and the AβPP-independent C100/C99 generation pathway is unconventionally activated. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates and crosses the T1 threshold. The disease recurs and proceeds uninterrupted. The treatment provides the temporary reprieve of probably several disease-free years.</p>
Full article ">Figure 24
<p><b>Overcoming limitations of the composite therapy for the prevention and treatment of unconventional AD: recurrent transient simultaneous administration of BACE activators and ISR inhibitors.</b> <span class="html-italic">i</span>Aβ: intraneuronal Aβ. <span class="html-italic">Blue lines</span>: levels of <span class="html-italic">i</span>Aβ in individual neurons. <span class="html-italic">T</span>1 <span class="html-italic">threshold</span>: cellular concentration of <span class="html-italic">i</span>Aβ that triggers the elicitation of the neuronal integrated stress response. <span class="html-italic">T</span>2 <span class="html-italic">threshold</span>: levels of <span class="html-italic">i</span>Aβ that trigger apoptosis or necroptosis of neuronal cells. <span class="html-italic">Red box</span>: apoptotic zone; within this zone, the neurons have either undergone apoptosis or are dead. <span class="html-italic">Yellow boxes</span>: duration of the concurrent administration of ISR inhibitors and BACE activators: (<b>A</b>) Recurrent implementation of the composite therapy in the prevention of unconventional AD. At the time of the initial round of the composite therapy, unconventional stressors have already occurred, the neuronal ISR is unconventionally elicited, and the AβPP-independent C100/C99 generation pathway is unconventionally activated. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates, but its levels are still below the T1 threshold. The transient administration of ISR inhibitors reverses the ISR state, disables the AβPP-independent C100/C99 generation, abolishes the influx of <span class="html-italic">i</span>Aβ produced independently of AβPP, restores the production of BACE enzymes, and reinstitutes their availability. The concurrent administration of BACE activators results in the efficient depletion of <span class="html-italic">i</span>Aβ. When both ISR inhibitors and BACE activators are withdrawn, the neuronal ISR is re-elicited due to the persistence of unconventional stressors, and the AβPP-independent C100/C99 generation pathway is unconventionally activated. The accumulation of <span class="html-italic">i</span>Aβ produced independently from AβPP resumes from a low baseline. Before it reaches the T1 threshold, the following rounds of the composite therapy are implemented repeatedly as needed. The T1 threshold is not crossed, and the disease does not occur as long as the composite therapy continues to be implemented. (<b>B</b>) Recurrent implementation of the composite therapy in the treatment of unconventional AD. The initial round of the composite therapy is administered to a symptomatic AD patient. By this time, the neuronal ISR has been unconventionally elicited and the AβPP-independent C100/C99 generation pathway unconventionally activated. <span class="html-italic">i</span>Aβ produced independently of AβPP rapidly accumulates and crosses the T1 threshold; the T2 threshold is reached in a fraction of neurons, and AD symptoms manifest. Transiently administered ISR inhibitors reverse the ISR state, restore the availability of BACE enzymes, and abrogate the influx of <span class="html-italic">i</span>Aβ produced independently of AβPP. Under these conditions, the concurrently administered BACE1 and/or BACE2 activators efficiently deplete <span class="html-italic">i</span>Aβ. Upon the conclusion of the treatment, the neuronal ISR is unconventionally re-elicited, and the AβPP-independent C100/C99 generation pathway is reactivated. The accumulation of <span class="html-italic">i</span>Aβ produced independently of AβPP resumes from a low baseline. Before it reaches the AD pathology-causing range, the following rounds of the composite therapy are administered repeatedly as required. The AD pathology-causing range of <span class="html-italic">i</span>Aβ concentrations is not reached, and AD symptoms do not recur for the duration of the treatment.</p>
Full article ">Figure 25
<p><b>C99 generated independently of AβPP as the driver of Alzheimer’s disease in its both conventional and unconventional forms. <span class="html-italic">i</span>Aβ: intraneuronal Aβ.</b> <span class="html-italic">AβPP</span>: Aβ protein precursor. <span class="html-italic">C99</span>: C-terminal fragment of AβPP produced either by the proteolysis of AβPP or independently of AβPP. <span class="html-italic">eIF2α:</span> eukaryotic translation initiation factor 2<span class="html-italic">α</span>. <span class="html-italic">PKR and HRI:</span> kinases capable of phosphorylating eIF2α when activated. <span class="html-italic">TNFα</span>: tumor necrosis factor α, potentially capable of activating PKR. <span class="html-italic">PACT</span>: PKR activator. <span class="html-italic">OMA1:</span> mitochondrial distress-activated mitochondrial protease. <span class="html-italic">DELE1:</span> substrate of OMA1; its cleavage leads to the activation of HRI. <span class="html-italic">ISR:</span> neuronal integrated stress response; it is elicited by phosphorylation of eIF2α and enables the production of components essential for the operation of the AβPP-independent <span class="html-italic">i</span>Aβ generation pathway. In conventional AD, the elicitation of the neuronal ISR is triggered via the activation of PKR and/or HRI kinases and phosphorylation of eIF2α, by AβPP-derived <span class="html-italic">i</span>Aβ accumulated over the T1 threshold. In unconventional AD, the elicitation of the neuronal ISR is triggered via the activation of one or more eIF2α kinases, by stressors other than AβPP-derived <span class="html-italic">i</span>Aβ. In both forms of AD, the neuronal ISR suppresses, in the context of the inhibition of the global cellular cap-dependent protein synthesis, the production of the components of the AβPP proteolytic pathway in general and of γ-secretase in particular. Simultaneously, it also enables, via the supply of essential components, the operation of the AβPP-independent C100/C99 generation pathway. The resulting C99, produced independently of AβPP, drives AD pathology, propagates the neuronal ISR, and perpetuates its own production.</p>
Full article ">
27 pages, 3932 KiB  
Article
Evaluation of the Anti-Amyloid and Anti-Inflammatory Properties of a Novel Vanadium(IV)–Curcumin Complex in Lipopolysaccharides-Stimulated Primary Rat Neuron-Microglia Mixed Cultures
by Georgios Katsipis, Sophia N. Lavrentiadou, George D. Geromichalos, Maria P. Tsantarliotou, Eleftherios Halevas, George Litsardakis and Anastasia A. Pantazaki
Int. J. Mol. Sci. 2025, 26(1), 282; https://doi.org/10.3390/ijms26010282 - 31 Dec 2024
Viewed by 275
Abstract
Lipopolysaccharides (LPS) are bacterial mediators of neuroinflammation that have been detected in close association with pathological protein aggregations of Alzheimer’s disease. LPS induce the release of cytokines by microglia and mediate the upregulation of inducible nitric oxide synthase (iNOS)—a mechanism also associated with [...] Read more.
Lipopolysaccharides (LPS) are bacterial mediators of neuroinflammation that have been detected in close association with pathological protein aggregations of Alzheimer’s disease. LPS induce the release of cytokines by microglia and mediate the upregulation of inducible nitric oxide synthase (iNOS)—a mechanism also associated with amyloidosis. Curcumin is a recognized natural medicine but has extremely low bioavailability. V-Cur, a novel hemocompatible Vanadium(IV)-curcumin complex with higher solubility and bioactivity than curcumin, is studied here. Co-cultures consisting of rat primary neurons and microglia were treated with LPS and/or curcumin or V-Cur. V-Cur disrupted LPS-induced overexpression of amyloid precursor protein (APP) and the in vitro aggregation of human insulin (HI), more effectively than curcumin. Cell stimulation with LPS also increased full-length, inactive, and total iNOS levels, and the inflammation markers IL-1β and TNF-α. Both curcumin and V-Cur alleviated these effects, with V-Cur reducing iNOS levels more than curcumin. Complementary insights into possible bioactivity mechanisms of both curcumin and V-Cur were provided by In silico molecular docking calculations on Aβ1-42, APP, Aβ fibrils, HI, and iNOS. This study renders curcumin-based compounds a promising anti-inflammatory intervention that may be proven a strong tool in the effort to mitigate neurodegenerative disease pathology and neuroinflammatory conditions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Levels of amyloid precursor protein (APP) in mixed cultures of primary neuron-microglia, in the absence or presence of 0.1, 1, or 10 μg/mL of LPS (<b>a</b>,<b>b</b>). Effect of LPS (1 μg/mL) in the presence or absence of 2 μΜ curcumin or V-Cur complex on APP levels (<b>c</b>,<b>d</b>). Analysis performed with Western blotting. Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was used to verify equal loading. The density of the blots was semi-quantified with ImageJ 1.54. Results are presented as fractional changes in comparison with the control sample and are the mean (±SD) of three independent biological experiments. One-way ANOVA was employed to compare untreated or LPS-treated samples. Statistical significance when compared with: * untreated sample (control); # LPS-treated sample; <span>$</span> curcumin-treated: * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001; #### <span class="html-italic">p</span> &lt; 0.0001; <span>$</span><span>$</span><span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 2
<p>Docking poses orientation of curcumin and V-Cur in the crystal structure of the Kunitz protease inhibitor domain (APPI) of APP (PDB accession number 1AAP). The target protein is illustrated as a semi-transparent cartoon and surface colored in yellow orange and chocolate (chains A and B, respectively), while curcumin and V-Cur molecules are rendered in ball-and-stick mode and colored according to atom type in light pink and violet purple C atoms, respectively. The ligand binding site of both molecules depicting the architecture of the binding interactions is also illustrated (in the upper part) with an additional depiction of selected contacting amino acid residues of the binding pocket rendered in line and colored according to the cartoon. Binding interactions are illustrated in light pink (for curcumin) and violet (for V-Cur). Heteroatom color code: V: grey, N: blue, and O: red. Molecular docking simulations of both ligands were performed individually. Hydrogen atoms are omitted for clarity. The final structure was ray-traced and illustrated with the aid of PyMol Molecular Graphics.</p>
Full article ">Figure 3
<p>In vitro fibrillation assay with insulin in the presence of several concentrations (0–100 μΜ) of either curcumin (■) or V-Cur (●). The insulin amyloid fibers formed in the absence or presence of either curcumin or V-Cur were semi-quantified by employing Thioflavin T fluorescence, with excitation at 450 nm and recording the emission spectrum at 490 nm. The results from three independent experiments are provided as mean normalized fibrillization rates (±SEM), setting the value of the control sample as 100%. Some error bars are not visible due to very small values (&lt;1%).</p>
Full article ">Figure 4
<p>Docking pose orientation of curcumin and V-Cur in the crystal structure of dimer (PDB ID 1GUJ) and hexamer (PDB ID 6GNQ) HΙ target proteins. In the hexameric structure of HI are also illustrated the six chain-stabilizing Zn<sup>2+</sup> ions, the co-crystallized meta-cresol (CRS, depicted in gold sticks), and some critical to self-assembly and aggregation resides of HI (represented in stick mode colored in orange). Both HI proteins’ structures are depicted as cartoon colored in wheat and firebrick for A and B chains, respectively. Curcumin and V-Cur are rendered in sphere representation colored according to atom type in light pink and violet purple, respectively. The two Zn ions co-crystallized in the hexameric structure are depicted in sphere representation in lemon color and are shown to be connected with polar contact with Nε2 of His10 in the three double chains (A and B). Heteroatom color code: V: grey, N: blue, and O: red. Molecular docking simulations of both ligands were performed individually. Hydrogen atoms are omitted for clarity. The final structure was ray-traced and illustrated with the aid of PyMol Molecular Graphics.</p>
Full article ">Figure 5
<p>Levels of (<b>a</b>) active inducible NO synthase (iNOS) (100 kDa), (<b>b</b>) inactive iNOS (50 and 75 kDa), and (<b>c</b>) total iNOS levels, after 24 h of treatment with LPS 1 μg/mL, in the presence or absence of 2 μΜ of curcumin or V-Cur complex, in mixed cultures of primary neurons-microglia. iNOS levels were determined by Western blotting (<b>d</b>). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was employed to verify equal loading. The density of the blots was semi-quantified with ImageJ 1.54. Results are presented as fractional changes in comparison with the control sample and are the mean (±SD) of three independent biological experiments. One-way ANOVA was employed to compare untreated or LPS-treated samples. Statistical significance when compared with: * untreated sample (control); # LPS-treated sample; <span>$</span> curcumin-treated: * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001; # <span class="html-italic">p</span> &lt; 0.05; ## <span class="html-italic">p</span> &lt; 0.01; ### <span class="html-italic">p</span> &lt; 0.001; #### <span class="html-italic">p</span> &lt; 0.0001; <span>$</span> <span class="html-italic">p</span> &lt; 0.05; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>Docking pose orientation of curcumin and V-Cur in the crystal structure of iNOS monomer enzyme (PDB accession number 4NOS). The target protein is illustrated as cartoon colored in the sand along with a semi-transparent surface colored in the dark sand. Curcumin and V-Cur molecules, as well as the co-crystallized iNOS inhibitor ethylisothiourea (ITU) are rendered in sphere mode and colored according to atom type in light pink, violet purple, and hot pink C atoms, respectively. The co-crystallized molecules heme (HEM) (iron protoporphyrin IX) and H2B superimposed with the docked molecules are rendered in stick representation and colored according to atom type in orange and yellow-orange C atoms, respectively. H4B, essential for the dimerization of the protein, is not shown since it is located farther down the binding cavity, near the dimerization interface. The target protein structure model in the lower panel, depicting in a close-up view of the binding cavity of the target enzyme the architecture of the binding interactions, is illustrated as a semi-transparent surface colored in dark sand with an additional depiction of selected contacting amino acid residues of the binding pocket highlighted in the molecular surface in smudge green (for V-Cur) and white (for curcumin). Binding interaction residues are labeled in white (for curcumin) and smudge green (for V-Cur). Heteroatom color code: V: grey, N: blue, and O: red. Molecular docking simulations of both ligands were performed individually. Hydrogen atoms are omitted for clarity. The final structure was ray-traced and illustrated with the aid of PyMol Molecular Graphics.</p>
Full article ">Figure 7
<p>Levels of (<b>a</b>) tumor necrosis factor-α (TNF-α), and (<b>b</b>) interleukin-1β (IL-1β), in mixed cultures of primary neurons-microglia after treatment with LPS (1 μg/mL) in the presence or absence of 2 μΜ curcumin or V-Cur. Cytokine levels were determined with Western blotting (<b>c</b>,<b>d</b>). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was employed to verify equal loading. The density of the blots was semi-quantified with ImageJ 1.54. Results are presented as fractional changes in comparison with the control sample and are the mean (± SD) of three independent biological experiments. One-way ANOVA was employed to compare untreated or LPS-treated samples. Statistical significance when compared with: * untreated sample (control); # LPS-treated sample; <span>$</span> curcumin-treated: ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001; # <span class="html-italic">p</span> &lt; 0.05; #### <span class="html-italic">p</span> &lt; 0.0001; <span>$</span><span>$</span> <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
21 pages, 3475 KiB  
Review
VDAC1: A Key Player in the Mitochondrial Landscape of Neurodegeneration
by Shirel Argueti-Ostrovsky, Shir Barel, Joy Kahn and Adrian Israelson
Biomolecules 2025, 15(1), 33; https://doi.org/10.3390/biom15010033 - 30 Dec 2024
Viewed by 338
Abstract
Voltage-Dependent Anion Channel 1 (VDAC1) is a mitochondrial outer membrane protein that plays a crucial role in regulating cellular energy metabolism and apoptosis by mediating the exchange of ions and metabolites between mitochondria and the cytosol. Mitochondrial dysfunction and oxidative stress are central [...] Read more.
Voltage-Dependent Anion Channel 1 (VDAC1) is a mitochondrial outer membrane protein that plays a crucial role in regulating cellular energy metabolism and apoptosis by mediating the exchange of ions and metabolites between mitochondria and the cytosol. Mitochondrial dysfunction and oxidative stress are central features of neurodegenerative diseases. The pivotal functions of VDAC1 in controlling mitochondrial membrane permeability, regulating calcium balance, and facilitating programmed cell death pathways, position it as a key determinant in the delicate balance between neuronal viability and degeneration. Accordingly, increasing evidence suggests that VDAC1 is implicated in the pathophysiology of neurodegenerative diseases, including Alzheimer’s disease (AD), Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), and others. This review summarizes the current findings on the contribution of VDAC1 to neurodegeneration, focusing on its interactions with disease-specific proteins, such as amyloid-β, α-synuclein, and mutant SOD1. By unraveling the complex involvement of VDAC1 in neurodegenerative processes, this review highlights potential avenues for future research and drug development aimed at alleviating mitochondrial-related neurodegeneration. Full article
Show Figures

Figure 1

Figure 1
<p>VDAC1 as a key modulator of mitochondrial metabolism, apoptosis, and quality control. The mitochondrion contains an outer and inner membrane, with an intermembrane space between them. The outer membrane includes VDAC1 channels that allow for the passage of ions (K<sup>+</sup>, Na<sup>+</sup>, Ca<sup>2+</sup>, Cl<sup>−</sup>) and metabolites (such as ATP, ADP, AMP, NADH, malate, citrate, glucose, pyruvate, succinate, and glutamate). Bcl-2 and HK are associated with VDAC1, regulating cellular apoptosis and glucose metabolism. Proteins such as PINK and Parkin are involved in mitochondrial quality control, specifically in mitophagy and apoptosis. VDAC1 serves as a substrate for Parkin-mediated ubiquitination where monoubiquitination inhibits apoptosis, while polyubiquitination promotes mitophagy. The mitochondrial permeability transition pore (mPTP) facilitates the release of Ca<sup>2+</sup> and cytochrome c, which are critical for initiating apoptosis. Within the inner membrane, complexes I-IV of the electron transport chain generate a proton gradient (H<sup>+</sup>) across the membrane. ATP synthase uses this gradient to synthesize ATP from ADP and inorganic phosphate (Pi). Mitochondrial DNA (mtDNA) is also shown, illustrating its role in encoding essential proteins for mitochondrial function.</p>
Full article ">Figure 2
<p>Proposed model for VDAC1 involvement in mitochondrial dysfunction in Alzheimer’s disease. In AD, the expression of VDAC1 is altered, and the channel conductance is increased. Aβ oligomers interact with VDAC1 at the plasma membrane through its N-terminal domain, leading to the formation of a VDAC1-Aβ heterooligomer that facilitates Aβ entry into the cell. Once in the cytosol, Aβ interacts with mitochondrial VDAC1, causing VDAC1 oligomerization [<a href="#B45-biomolecules-15-00033" class="html-bibr">45</a>]. Moreover, VDAC1 directly interacts with phosphorylated tau (pTau), further contributing to mitochondrial dysfunction. Finally, increased activity of GSK3β promotes its translocation to the mitochondria, and the phosphorylation of VDAC1 at threonine 51 (Thr51). This phosphorylation causes the detachment of HK from VDAC1, triggering cytochrome c release and promoting apoptotic cell death.</p>
Full article ">Figure 3
<p>Proposed model for VDAC1 involvement in mitochondrial dysfunction in Parkinson’s disease. In PD, VDAC1 altered expression and its interactions with α-synuclein (α-syn) disrupt mitochondrial function. This interaction enables α-synuclein translocation and accumulation to the mitochondria, modulating VDAC1 permeability and increasing selective calcium flux through the VDAC1 channel, leading to mitochondrial calcium overload, dysfunction, and subsequent cell death. Additionally, VDAC1 is involved in the IP3R3-Grp75-VDAC1 complex at mitochondria-associated membranes (MAM), which mediates calcium transfer from the endoplasmic reticulum (ER) to mitochondria. This complex is destabilized by the loss of DJ-1, impairing ER–mitochondria calcium homeostasis. Mutations in Parkin inhibit VDAC1 monoubiquitination, leading to increased mitochondrial calcium influx and apoptosis, reflecting PD pathology. HK detachment from VDAC1 results in cytochrome c release, amplifying apoptotic signaling. In addition, defects in complex I of the electron transport chain are a major cause of mitochondrial malfunction in PD.</p>
Full article ">Figure 4
<p>Proposed model for VDAC1 involvement in mitochondrial dysfunction in ALS. In ALS, mutant SOD1 interacts with VDAC1 through its N-terminal domain, leading to the detachment of HK1 from the channel. In addition, the association between mutant SOD1 and Bcl-2 disrupts the association between Bcl-2 and VDAC1. Moreover, VDAC1 oligomerization levels are elevated, channel conductance is reduced, and VDAC1 undergoes specific Asn<sup>207</sup> deamidation and Met<sup>155</sup> oxidation. In TDP-43-related ALS, TDP-43 mislocalizes into the mitochondria and interacts with VDAC1, resulting in aberrant mtDNA release to the cytosol.</p>
Full article ">
18 pages, 3842 KiB  
Article
Co-Localized in Amyloid Plaques Cathepsin B as a Source of Peptide Analogs Potential Drug Candidates for Alzheimer’s Disease
by Marilena K. Theodoropoulou, Konstantina D. Vraila, Nikos C. Papandreou, Georgia I. Nasi and Vassiliki A. Iconomidou
Biomolecules 2025, 15(1), 28; https://doi.org/10.3390/biom15010028 - 30 Dec 2024
Viewed by 317
Abstract
Alzheimer’s disease (AD) is a complex neurodegenerative disorder characterized by extracellular amyloid plaques, predominantly consisting of amyloid-β (Aβ) peptides. The oligomeric form of Aβ is acknowledged as the most neurotoxic, propelling the pathological progression of AD. Interestingly, besides A [...] Read more.
Alzheimer’s disease (AD) is a complex neurodegenerative disorder characterized by extracellular amyloid plaques, predominantly consisting of amyloid-β (Aβ) peptides. The oligomeric form of Aβ is acknowledged as the most neurotoxic, propelling the pathological progression of AD. Interestingly, besides Aβ, other proteins are co-localized within amyloid plaques. Peptide analogs corresponding to the “aggregation-prone” regions (APRs) of these proteins could exhibit high-affinity binding to Aβ and significant inhibitory potential against the Aβ oligomerization process. The peptide analogs of co-localized protease, Cathepsin B, may act as such potent inhibitors. In silico studies on the complexes of the oligomeric state of Aβ and Cathepsin B peptide analogs were performed utilizing molecular docking and molecular dynamics simulations, revealing that these analogs disrupt the β-sheet-rich core of Aβ oligomers, a critical structural feature of their stability. Of the four peptide analogs evaluated, two demonstrated considerable potential by effectively destabilizing oligomers while maintaining low self-aggregation propensity, i.e., a crucial consideration for therapeutic safety. These findings point out the potential of APR-derived peptide analogs from co-localized proteins as innovative agents against AD, paving the way for further exploration in peptide-based therapeutic development. Full article
(This article belongs to the Special Issue Amyloid-Beta and Alzheimer’s Disease)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the methods employed in the present study, alongside the materials and programs deployed.</p>
Full article ">Figure 2
<p>Graphical representation of the predicted “aggregation-prone” regions on the functional dimer of Cathepsin B. The CathB1 peptide analog is colored orange, the CathB1a peptide analog is yellow, the CathB2 peptide analog is forest green, and the CathB2a peptide analog is purple.</p>
Full article ">Figure 3
<p>Results of molecular dynamic simulations of A<span class="html-italic">β</span><sub>42</sub> oligomer with the CathB1 peptide analog. (<b>A</b>) The number of residues assigned to the beta-sheet (blue), 3(10)-helix (orange), and a-helix (grey) of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the simulation, according to DSSP calculations. (<b>B</b>) The RMSD of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the course of the simulation. (<b>C</b>) The RMSF calculation for the residues of chain A (light blue), chain B (orange), chain C (grey), chain D (yellow), and chain E (dark blue) of the A<span class="html-italic">β</span><sub>42</sub> oligomer. (<b>D</b>) The number of hydrogen bonds formed between the chains of the A<span class="html-italic">β</span><sub>42</sub> oligomer and more particularly between chains A and B (blue), chains B and C (orange), chains C and D (grey), and chains D and E (yellow).</p>
Full article ">Figure 4
<p>The results of molecular dynamic simulations of A<span class="html-italic">β</span><sub>42</sub> oligomer with CathB2 peptide analog. (<b>A</b>) The number of residues assigned to beta-sheet (blue), 3(10)-helix (orange), and a-helix (grey) of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the simulation, according to DSSP calculations. (<b>B</b>) The RMSD of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the course of the simulation. (<b>C</b>) The RMSF calculation for the residues of chain A (light blue), chain B (orange), chain C (grey), chain D (yellow), and chain E (blue) of the A<span class="html-italic">β</span><sub>42</sub> oligomer. (<b>D</b>) The number of hydrogen bonds formed between the chains of A<span class="html-italic">β</span><sub>42</sub> oligomer and, more particularly, between chains A and B (blue), chains B and C (orange), chains C and D (grey), chains D and E (yellow).</p>
Full article ">Figure 5
<p>The results of molecular dynamic simulations of A<span class="html-italic">β</span><sub>42</sub> oligomer with CathB1a peptide analog. (<b>A</b>) The number of residues assigned to beta-sheet (blue), 3(10)-helix (orange), and a-helix (grey) of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the simulation, according to DSSP calculations. (<b>B</b>) The RMSD of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the course of the simulation. (<b>C</b>) The RMSF calculation for the residues of chain A (light blue), chain B (orange), chain C (grey), chain D (yellow), and chain E (blue) of the A<span class="html-italic">β</span><sub>42</sub> oligomer. (<b>D</b>) The number of hydrogen bonds formed between the chains of A<span class="html-italic">β</span><sub>42</sub> oligomer and more particularly between chains A and B (blue), chains B and C (orange), chains C and D (grey), and chains D and E (yellow).</p>
Full article ">Figure 6
<p>The results of molecular dynamic simulations of A<span class="html-italic">β</span><sub>42</sub> oligomer with CathB2a peptide analog. (<b>A</b>) The number of residues assigned to beta-sheet (blue), 3(10)-helix (orange), and a-helix (grey) of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the simulation, according to DSSP calculations. (<b>B</b>) The RMSD of the A<span class="html-italic">β</span><sub>42</sub> oligomer during the course of the simulation. (<b>C</b>) The RMSF calculation for the residues of chain A (light blue), chain B (orange), chain C (grey), chain D (yellow), and chain E (blue) of the A<span class="html-italic">β</span><sub>42</sub> oligomer. (<b>D</b>) The number of hydrogen bonds formed between the chains of A<span class="html-italic">β</span><sub>42</sub> oligomer and more particularly between chains A and B (blue), chains B and C (orange), chains C and D (grey), and chains D and E (yellow).</p>
Full article ">Figure 7
<p>PMF profiles for peptide–A<span class="html-italic">β</span> oligomer complexes. Subfigures (<b>A</b>–<b>D</b>) depict the energy fluctuations (kcal/mol) relative to the reaction coordinate (ξ) for peptide analogs CathB1 (<b>A</b>), CathB2 (<b>B</b>), CathB1a (<b>C</b>), and CathB2a (<b>D</b>).</p>
Full article ">Figure 8
<p>MD simulations of intermolecular interactions between CathB1a and CathB2a peptide analogs. The number of hydrogen bonds formed between the peptide analogs were calculated: (<b>A</b>) CathB1a; (<b>B</b>) CathB2a.</p>
Full article ">
28 pages, 2241 KiB  
Review
Novel Role of Pin1-Cis P-Tau-ApoE Axis in the Pathogenesis of Preeclampsia and Its Connection with Dementia
by Emmanuel Amabebe, Zheping Huang, Sukanta Jash, Balaji Krishnan, Shibin Cheng, Akitoshi Nakashima, Yitong Li, Zhixong Li, Ruizhi Wang, Ramkumar Menon, Xiao Zhen Zhou, Kun Ping Lu and Surendra Sharma
Biomedicines 2025, 13(1), 29; https://doi.org/10.3390/biomedicines13010029 - 26 Dec 2024
Viewed by 405
Abstract
Preeclampsia (preE) is a severe multisystem hypertensive syndrome of pregnancy associated with ischemia/hypoxia, angiogenic imbalance, apolipoprotein E (ApoE)-mediated dyslipidemia, placental insufficiency, and inflammation at the maternal–fetal interface. Our recent data further suggest that preE is associated with impaired autophagy, vascular dysfunction, and proteinopathy/tauopathy [...] Read more.
Preeclampsia (preE) is a severe multisystem hypertensive syndrome of pregnancy associated with ischemia/hypoxia, angiogenic imbalance, apolipoprotein E (ApoE)-mediated dyslipidemia, placental insufficiency, and inflammation at the maternal–fetal interface. Our recent data further suggest that preE is associated with impaired autophagy, vascular dysfunction, and proteinopathy/tauopathy disorder, similar to neurodegenerative diseases such as Alzheimer’s disease (AD), including the presence of the cis stereo-isoform of phosphorylated tau (cis P-tau), amyloid-β, and transthyretin in the placenta and circulation. This review provides an overview of the factors that may lead to the induction and accumulation of cis P-tau-like proteins by focusing on the inactivation of peptidyl-prolyl cis–trans isomerase (Pin1) that catalyzes the cis to trans isomerization of P-tau. We also highlighted the novel role of the Pin1-cis P-tau-ApoE axis in the development of preE, and propagation of cis P-tau-mediated abnormal protein aggregation (tauopathy) from the placenta to cerebral tissues later in life, leading to neurodegenerative conditions. In the case of preE, proteinopathy/tauopathy may interrupt trophoblast differentiation and induce cell death, similar to the events occurring in neurons. These events may eventually damage the endothelium and cause systemic features of disorders such as preE. Despite impressive research and therapeutic advances in both fields of preE and neurodegenerative diseases, further investigation of Pin1-cis P-tau and ApoE-related mechanistic underpinnings may unravel novel therapeutic options, and new transcriptional and proteomic markers. This review will also cover genetic polymorphisms in the ApoE alleles leading to dyslipidemia induction that may regulate the pathways causing preE or dementia-like features in the reproductive age or later in life, respectively. Full article
(This article belongs to the Special Issue Pathogenesis and Treatment of Preeclampsia)
Show Figures

Figure 1

Figure 1
<p>I. Inactivation of the Pin1-cisP-tau axis induces preeclampsia features. Exposure of trophoblast cells to hypoxic stress or serum from preeclampsia patients inhibits Pin1 by oxidation or phosphorylation, thereby facilitating cis P-tau aggregation. The protein aggregation may also be secondary to impaired autophagic clearance induced by hypoxia or preeclampsia serum. Elevated cis P-tau aggregates inhibit the interaction between EVT and spiral artery endothelium, leading to shallow EVT invasion and poor vascular remodeling that precipitate preeclampsia features. Hypoxia also induces the production of sFlt-1 and sEng in trophoblasts mediated by or independent of cis P-tau aggregation. sFlt-1 and sEng can cause vascular endothelial dysfunction that ultimately results in the preeclampsia syndrome. Importantly, cis mAb inhibits the accumulation of cis P-tau, thereby re-storing normal placental development and function. DAPK1, death associated protein kinase 1; EVT, extravillous trophoblast; FGR, fetal growth restriction; HELLP, hemolysis, elevated liver enzymes, and low platelet count; Pin1, peptidyl-prolyl cis–trans isomerase; pS289-DAPK1, DAPK1 phosphorylated at serine residue 289; sEng, soluble endoglin; sFlt-1, fms-like tyrosine kinase 1. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 2
<p>Transcription factor EB mediates placental development. (<b>a</b>) TFEB-mediated autophagy–lysosomal dysfunction contributes to protein (cis P-tau) aggregation associated with preeclampsia. Inhibition of nuclear translation or transcriptional activity of TFEB inhibits autophagy and disrupt lysosomal function through reduced protein expression and function. This leads to increased p62 and accumulation of toxic protein aggregates, including cis P-tau, which can leak into circulation (cistauosis). Increased accumulation of p62 in trophoblast cells leads to reduced placental growth factor (PlGF). Moreover, increased serum sEng enhances lysosomal dysfunction. The resultant effect of all these alterations is increased apoptosis, reduced EVT invasion, and inadequate spiral artery remodeling, ultimately inducing preeclampsia features. Phosphorylation (by mTORC1) and acetylation (by GCN5) of TFEB inhibit its nuclear translocation and transcriptional activity, respectively. TFEB nuclear export is also promoted by activated chromosome region maintenance 1 (CRM1). Trehalose is a mTORC1-independent autophagy promoter. Trehalose and lactotrehalose promote autophagy by increasing the nuclear translocation of TFEB. (<b>b</b>) Autophagy-independent TFEB-mediated maintenance of placental function. Deficiency or reduced nuclear translocation of TFEB impairs syncytiotrophoblast (STB) formation, placental vascular construction, and hormone production, including estradiol through reduced activation of the fusogens <span class="html-italic">ERVFRD-1</span> and <span class="html-italic">ERVW-1</span>. These factors can be independently or in synergy with dysfunctional autophagy–lysosomal biogenesis machinery to induce features of preeclampsia. <span class="html-italic">CRM1</span>, chromosome region maintenance 1; <span class="html-italic">ERVFRD-1</span>, endogenous retrovirus group FRD member 1 (Syncytin-2); <span class="html-italic">ERVW-1</span>, endogenous retrovirus group W member 1 (Syncytin-1); <span class="html-italic">FGR</span>, fetal growth restriction; <span class="html-italic">GCN5</span>, general control non-repressed protein 5; <span class="html-italic">HELLP</span>, hemolysis, elevated liver enzymes, and low platelet count; <span class="html-italic">mTORC1</span>, mechanistic target of rapamycin complex 1; <span class="html-italic">TFEB</span>, transcription factor of the E box. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 3
<p>The mechanistic link between preeclampsia and dementia. Both preeclampsia and dementia are mediated by <span class="html-italic">endothelial dysfunction</span> induced by the accumulation and spread of cis P-tau aggregates in placental and brain tissues, respectively. cis P-tau accumulates due to inactivated Pin1 and impaired autophagy and lysosomal dysfunction induced by hypoxia secondary to hypoperfusion. Cistauosis due to preeclampsia-related leakage of accumulated placental cis P-tau into systemic circulation can also deliver cis P-tau to brain neurons in the long run, leading to cognitive impairment and dementia. <span class="html-italic">Pin1</span>, peptidyl-prolyl cis–trans isomerase; <span class="html-italic">P-tau</span>, phosphorylated tau. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">Figure 4
<p>Mechanisms of ApoE-mediated dyslipidemia, vascular dysfunction, and tauopathy. The accumulation of cis P-tau aggregates seen in brain and placental tissues of Alzheimer’s disease (AD) and preeclampsia patients, respectively, could be induced by hypoxia-associated impaired autophagy and Pin1 activation precipitated by vascular damage due to dyslipidemia. This sequence of pathologic events is seen in ApoE4 carriers, while ApoE2 seems to confer some protection by preventing vascular damage through the reduction of oxidative stress marker (MDA), LDL and increasing HDL. <span class="html-italic">ADRD</span>, Alzheimer’s disease-related disorders; <span class="html-italic">ApoE</span>, apolipoprotein E; <span class="html-italic">HDL</span>, high-density lipoprotein; <span class="html-italic">LDL</span>, low-density lipoprotein; <span class="html-italic">MDA</span>, malondialdehyde; <span class="html-italic">P-tau</span>, phosphorylated tau; <span class="html-italic">Pin1</span>, peptidyl-prolyl cis–trans isomerase; <span class="html-italic">VLDL</span>, very low-density lipoprotein. Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>.</p>
Full article ">
16 pages, 2562 KiB  
Article
Rosemarinic Acid-Induced Destabilization of Aβ Peptides: Insights from Molecular Dynamics Simulations
by Liang Zhao, Weiye Jiang, Zehui Zhu, Fei Pan, Xin Xing, Feng Zhou and Lei Zhao
Foods 2024, 13(24), 4170; https://doi.org/10.3390/foods13244170 - 23 Dec 2024
Viewed by 473
Abstract
Alzheimer’s disease (AD) is a neurodegenerative disorder marked by the progressive accumulation of amyloid-β (Aβ) plaques and tau protein tangles in the brain. These pathological aggregates interfere with neuronal function, leading to the disruption of cognitive processes, particularly memory. The deposition of Aβ [...] Read more.
Alzheimer’s disease (AD) is a neurodegenerative disorder marked by the progressive accumulation of amyloid-β (Aβ) plaques and tau protein tangles in the brain. These pathological aggregates interfere with neuronal function, leading to the disruption of cognitive processes, particularly memory. The deposition of Aβ forms senile plaques, while tau protein, in its hyperphosphorylated state, forms neurofibrillary tangles, both of which contribute to the underlying neurodegeneration observed in AD. Rosmarinic acid (RosA), a natural compound found in plants such as Rosmarinus officinalis, is known for its antioxidant, anti-inflammatory, and antimicrobial properties. Due to its ability to cross the blood–brain barrier, RosA holds promise as a nutritional supplement that may support brain health. In this study, molecular dynamics (MD) simulations were used to investigate the impact of RosA on the structural stability of Aβ peptides. The results indicated that the addition of RosA increased the instability of Aβ, as evidenced by an increase in the Root Mean Square Deviation (RMSD), a decrease in the Radius of Gyration (Rg), and an expansion of the Solvent Accessible Surface Area (SASA). This destabilization is primarily attributed to the disruption of native hydrogen bonds and hydrophobic interactions in the presence of two RosA molecules. The free energy landscape (FEL) analysis and MM-PBSA (Poisson-Boltzmann Surface Area Mechanics) results further support the notion that RosA can effectively bind to the hydrophobic pocket of the protein, highlighting its potential as a nutritional component that may contribute to maintaining brain health and function. Full article
(This article belongs to the Special Issue Development and Evaluation of Novel Functional Foods)
Show Figures

Figure 1

Figure 1
<p>Molecular docking analysis: best-chosen conformation of the binding sites. (<b>A</b>) Stabilized 3D conformation of RosA bound to Aβ peptide; (<b>B</b>) hydrophobic pocket between RosA and Aβ peptide binding sites; (<b>C</b>) 2D view of the interaction of RosA<sub>1</sub> and (<b>D</b>) RosA<sub>2</sub> binding sites; (<b>E</b>) 3D view of RosA<sub>1</sub>; and (<b>F</b>) the binding site interactions of the RosA<sub>2</sub> peptide are shown, with the ligand depicted in gray and hydrogen bonds to the protein residues highlighted in green.</p>
Full article ">Figure 2
<p>(<b>A</b>) Cartoon representation of Aβ peptide; (<b>B</b>) the 2D chemical structure of RosA.</p>
Full article ">Figure 3
<p>(<b>A</b>) Molecular trajectories of Aβ peptide systems during 500 ns MD simulations; (<b>B</b>) molecular trajectories of the Aβ-RosA systems during 500 ns MD simulations.</p>
Full article ">Figure 4
<p>Basic results of MD simulation based on the molecular docking structure. (<b>A</b>) The RMSD values of systems; (<b>B</b>) the Rg values of systems; (<b>C</b>) the SASA values of systems; (<b>D</b>) the RMSF values of systems. The horizontal axes labeled Ch_A, Ch_B, Ch_C, Ch_D, Ch_E, and Ch_F correspond to the individual chains A through F of the Aβ peptide.</p>
Full article ">Figure 5
<p>DSSP-based secondary structure time evolution over 500 ns. (<b>A</b>) The secondary structure time evolution of Aβ peptide; (<b>B</b>) the secondary structure time evolution of Aβ-RosA system.</p>
Full article ">Figure 6
<p>Binding energy between RosA and Aβ peptide calculated using MM-PBSA, along with the breakdown of individual energy components. All free energy values are given in units of kcal/mol.</p>
Full article ">Figure 7
<p>FEL results. (<b>A</b>) The lowest energy conformation diagram of the Aβ protein–RosA<sub>1</sub>; (<b>B</b>) the lowest energy conformation diagram of the Aβ peptide–RosA<sub>2</sub>; (<b>C</b>) 3D view of RosA<sub>1</sub>; (<b>D</b>) RosA<sub>2</sub> protein binding site interactions, where the ligand is indicated in gray and the hydrogen bonding with the protein amino acid is indicated in green; (<b>E</b>) 2D free energy landscape of RosA<sub>1</sub>; (<b>F</b>) 2D free energy landscape of RosA<sub>2</sub>; (<b>G</b>) 3D free energy landscape of RosA<sub>1</sub>; (<b>H</b>) 3D free energy landscape of RosA<sub>2</sub>.</p>
Full article ">
15 pages, 3307 KiB  
Article
Exposure to Cadmium and Other Trace Elements Among Individuals with Mild Cognitive Impairment
by Teresa Urbano, Marco Vinceti, Chiara Carbone, Lauren A. Wise, Marcella Malavolti, Manuela Tondelli, Roberta Bedin, Giulia Vinceti, Alessandro Marti, Annalisa Chiari, Giovanna Zamboni, Bernhard Michalke and Tommaso Filippini
Toxics 2024, 12(12), 933; https://doi.org/10.3390/toxics12120933 - 22 Dec 2024
Viewed by 528
Abstract
Background: A limited number of studies have investigated the role of environmental chemicals in the etiology of mild cognitive impairment (MCI). We performed a cross-sectional study of the association between exposure to selected trace elements and the biomarkers of cognitive decline. Methods: During [...] Read more.
Background: A limited number of studies have investigated the role of environmental chemicals in the etiology of mild cognitive impairment (MCI). We performed a cross-sectional study of the association between exposure to selected trace elements and the biomarkers of cognitive decline. Methods: During 2019–2021, we recruited 128 newly diagnosed patients with MCI from two Neurology Clinics in Northern Italy, i.e., Modena and Reggio Emilia. At baseline, we measured serum and cerebrospinal fluid (CSF) concentrations of cadmium, copper, iron, manganese, and zinc using inductively coupled plasma mass spectrometry. With immuno-enzymatic assays, we estimated concentrations of β-amyloid 1-40, β-amyloid 1-42, Total Tau and phosphorylated Tau181 proteins, neurofilament light chain (NfL), and the mini-mental state examination (MMSE) to assess cognitive status. We used spline regression to explore the shape of the association between exposure and each endpoint, adjusted for age at diagnosis, educational attainment, MMSE, and sex. Results: In analyses between the serum and CSF concentrations of trace metals, we found monotonic positive correlations between copper and zinc, while an inverse association was observed for cadmium. Serum cadmium concentrations were inversely associated with amyloid ratio and positively associated with Tau proteins. Serum iron concentrations showed the opposite trend, while copper, manganese, and zinc displayed heterogeneous non-linear associations with amyloid ratio and Tau biomarkers. Regarding CSF exposure biomarkers, only cadmium consistently showed an inverse association with amyloid ratio, while iron was positively associated with Tau. Cadmium concentrations in CSF were not appreciably associated with serum NfL levels, while we observed an inverted U-shaped association with CSF NfL, similar to that observed for copper. In CSF, zinc was the only trace element positively associated with NfL at high concentrations. Conclusions: In this cross-sectional study, high serum cadmium concentrations were associated with selected biomarkers of cognitive impairment. Findings for the other trace elements were difficult to interpret, showing complex and inconsistent associations with the neurodegenerative endpoints examined. Full article
(This article belongs to the Special Issue Cadmium and Trace Elements Toxicity)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study flowchart. Abbreviations: CSF, cerebrospinal fluid; SCD, subjective cognitive decline (SCD).</p>
Full article ">Figure 2
<p>Violin plots distribution of trace element concentrations in serum and cerebrospinal fluid (CSF) according to sex (M, males; F, females). MCI, serum n = 89; CSF n = 45. Abbreviations: Cd, cadmium; Cu, copper; Fe, iron; MCI, mild cognitive impairment; Mn, manganese; NfL, neurofilament light chain; SCD, subjective cognitive decline; Zn, zinc.</p>
Full article ">Figure 3
<p>Spline regression analysis of the association between trace element concentration in serum and cerebrospinal fluid among patients with mild cognitive impairment. The solid line indicates the multivariable analysis; the shaded area represents upper and lower confidence interval limits. The dashed line represents association assuming linearity. Diamonds represent individual observations (n = 45). Abbreviations: Cd, cadmium; CSF, cerebrospinal fluid; Cu, copper; Fe, iron; Mn, manganese; NfL, neurofilament light chain; Zn, zinc.</p>
Full article ">Figure 4
<p>Spline regression analysis of the association between trace element concentration in serum (dark blue) and cerebrospinal fluid (CSF-light blue) with serum (<b>A</b>) and CSF neurofilament light (NfL) concentrations (<b>B</b>) among patients with mild cognitive impairment. The solid line indicates the multivariable analysis; the shaded area represents the upper and lower confidence interval limits. The dashed line represents the association assuming linearity. Diamonds represent individual observations (serum n = 89; CSF n = 45). Abbreviations: Cd, cadmium; Cu, copper; Fe, iron; MMSE, mini-mental state examination; Mn, manganese; Zn, zinc.</p>
Full article ">Figure 5
<p>Spline regression analysis of the association between trace element concentration in serum (dark green) and cerebrospinal fluid (CSF—light green) with CSF concentration of amyloid ratio (<b>A</b>), Total Tau (<b>B</b>) and phosphorylated Tau (p-Tau181) protein (<b>C</b>) among patients with mild cognitive impairment. The solid line indicates the multivariable analysis; the shaded area represents the upper and lower confidence interval limits. The dashed line represents the association assuming linearity. Diamonds represent individual observations (serum n = 89; CSF n = 45). Red lines represent laboratory cut-offs (&gt;0.069 for amyloid ratio; &lt;400 pg/mL for Total Tau; &lt;56.5 pg/mL for p-Tau181). Abbreviations: Cd, cadmium; Cu, copper; Fe, iron; MMSE, mini-mental state examination; Mn, manganese; Zn, zinc.</p>
Full article ">
18 pages, 6538 KiB  
Article
Yeast Glucan Remodeling Protein Bgl2p: Amyloid Properties and the Mode of Attachment in Cell Wall
by Nikita A. Motorin, Gennady I. Makarov, Valentina V. Rekstina, Evgeniy G. Evtushenko, Fanis A. Sabirzyanov, Rustam H. Ziganshin, Alexey K. Shaytan and Tatyana S. Kalebina
Int. J. Mol. Sci. 2024, 25(24), 13703; https://doi.org/10.3390/ijms252413703 - 22 Dec 2024
Viewed by 349
Abstract
Bgl2p is a major, conservative, constitutive glucanosyltransglycosylase of the yeast cell wall (CW) with amyloid amino acid sequences, strongly non-covalently anchored in CW, but is able to leave it. In the environment, Bgl2p can form fibrils and/or participate in biofilm formation. Despite a [...] Read more.
Bgl2p is a major, conservative, constitutive glucanosyltransglycosylase of the yeast cell wall (CW) with amyloid amino acid sequences, strongly non-covalently anchored in CW, but is able to leave it. In the environment, Bgl2p can form fibrils and/or participate in biofilm formation. Despite a long study, the question of how Bgl2p is anchored in CW remains unclear. Earlier, it was demonstrated that Bgl2p lost the ability to attach in CW and to fibrillate after the deletion of nine amino acids in its C-terminal region (CTR). Here, we demonstrated that a Bgl2p anchoring is weakened by substitution Glu-233/Ala in the active center. Using AlphaFold and molecular modeling approach, we demonstrated the role of CTR on Bgl2p attachment and supposed the conformational possibilities determined by the presence or absence of an intramolecular disulfide bond, forming by Cys-310, leading to accessibility of amyloid sequence and β-turns localized in CTR of Bgl2p for protein interactions. We hypothesized the mode of Bgl2p attachment in CW. Using atomic force microscopy, we investigated fibrillar structures formed by peptide V187MANAFSYWQ196 and suggested that it can serve as a factor leading to the induction of amyloid formation during interaction of Bgl2p with other proteins and is of medical interest being located close to the surface of the molecule. Full article
(This article belongs to the Special Issue 25th Anniversary of IJMS: Advances in Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Potential energies of globules and entropies of its torsion angles for models of Bgl2p from <span class="html-italic">Saccharomyces cerevisiae</span> (red) and homologous glycosidases from the GH17 family (blue). Points show average values and error bars show root mean square deviations. Labels represent instruments applied to the generation of the Bgl2 model or PDB ID for structures of homologous glycosidases.</p>
Full article ">Figure 2
<p>The structure of Bgl2p and position of strong amyloidogenic sites F<sub>83</sub>TIFVGV<sub>89</sub> (Site I) are highlighted in coral, N<sub>190</sub>AFS<sub>193</sub> (Site II)—in purple and G<sub>268</sub>VNVIVFEA<sub>276</sub> (Site III)—in olive. Primary structure of Bgl2p with designations of the secondary structure elements (<b>A</b>): α-helices underlined as helices, β-strands underlined as arrows. Tertiary Bgl2p structure (<b>B</b>) and the part of a molecule with sites I, II and III (<b>C</b>). The black dashed lines show hydrogen bonds between amyloidogenic sites and the rest of the Bgl2p globule, shown by green tubes and ribbons. Oxygen, nitrogen and hydrogen atoms are shown as red, blue and grey sticks, correspondingly.</p>
Full article ">Figure 3
<p>Representative AFM images of V<sub>187</sub>MANAFSYWQ<sub>196</sub> peptide on Ultraflat polystyrene support. (<b>A</b>) Sample from the bulk liquid. (<b>B</b>,<b>C</b>) The sample was prepared by Langmuir–Schaefer and transferred from the liquid-air interface. Panel B illustrates the complete monolayer of short fibrils present on the liquid-air interface with rare fibrils on top (white). The inset contains 2× enlarged fragment of short fibrils monolayer for visual clarity. Panel C hows bundles of longer fibrils of various morphology on top of this monolayer. (<b>D</b>) The scheme of prepared samples revealed the localization of fibrils.</p>
Full article ">Figure 4
<p>Conformation of the C-terminal region of the Bgl2p structure model. The L<sub>308</sub>DCD<sub>311</sub> residues of the C-terminal region, forming unstable β-turn, are shown by a yellow tube. Phe-298 and Leu-304 residues, forming a stable β-bridge, are shown by violet rods, while Asp-302, Lys-305 and Tyr-306, forming stable β-turn, are shown by deepteal rods. The black dashed lines show hydrogen bonds. The laying of the rest of the globule is shown by green tubes and ribbons. Oxygen, nitrogen and hydrogen atoms are shown as red, blue and grey sticks, correspondingly.</p>
Full article ">Figure 5
<p>The impact of deletion of nine amino acid residues of Bgl2p from the C-terminus on the conformational stability of certain regions of the Bgl2p molecule. RMSF profiles in the C-terminal region along the sequence for Bgl2p and Bgl2pΔC<sub>305–313</sub> are shown with designations of the secondary structure elements: α-helices underlined as helices, β-strands underlined as arrows. Three curves for each system correspond to three independent calculations. The protein sites with mobility changes reproducibly are highlighted in color: region 192–211—orange, region 238–248—blue, region 255–260—purple, region 299–304—yellow.</p>
Full article ">Figure 6
<p>An illustration of the conformational mobility obtained during the MD calculations of (<b>A</b>) Bgl2p and (<b>B</b>) Bgl2pΔC<sub>305–313</sub>. The color scheme of the sites, the mobility of which changes with the deletion of the C-terminal region, corresponds to <a href="#ijms-25-13703-f005" class="html-fig">Figure 5</a>; the amino acid residues of the active center are marked in green (Glu-233—light green, Glu-124—dark green). An overlay of several frames from the MD trajectory is presented to illustrate conformational mobility/variability.</p>
Full article ">Figure 7
<p>The conformations of the loop forming by aa 238–248 affect the geometry of the Bgl2p catalytic gap. (<b>A</b>) Bgl2p structure. (<b>B</b>) Bgl2pΔC<sub>305–313</sub> structure. Loop 238–248 are depicted in cyan. Active site: Glu-233—light green, Glu-124—dark green. The dotted line indicates the catalytic gap. The red circle indicates the overlapping zone of the catalytic gap.</p>
Full article ">Figure 8
<p>Western blot stained with antibodies to Bgl2 of samples obtained from <span class="html-italic">Saccharomyces cerevisiae</span> cell walls of E233 control and E233A mutant strains, untreated (-) or pretreated (+) with 1% SDS, and then incubated with 0.1 M Tris. The remaining protein was extracted from CW into a Laemmli buffer with β-mercaptoethanol after Tris incubation (lanes 1). Proteins extracted in Tris (lanes 2). All extracts were equalized by the optical density of untreated CW.</p>
Full article ">Figure 9
<p>The conformational mobility of the C-terminus (marked in green) of Bgl2p structures, non-mutated (gray) and with C310A substitution (coral). RMSF profiles of three trajectories for both structures.</p>
Full article ">
27 pages, 2428 KiB  
Review
The Emerging Role of PCSK9 in the Pathogenesis of Alzheimer’s Disease: A Possible Target for the Disease Treatment
by Gabriella Testa, Serena Giannelli, Erica Staurenghi, Rebecca Cecci, Lucrezia Floro, Paola Gamba, Barbara Sottero and Gabriella Leonarduzzi
Int. J. Mol. Sci. 2024, 25(24), 13637; https://doi.org/10.3390/ijms252413637 - 20 Dec 2024
Viewed by 499
Abstract
Alzheimer’s disease (AD) is a multifactorial neurodegenerative disease mainly caused by β-amyloid (Aβ) accumulation in the brain. Among the several factors that may concur to AD development, elevated cholesterol levels and brain cholesterol dyshomeostasis have been recognized to play a relevant role. Proprotein [...] Read more.
Alzheimer’s disease (AD) is a multifactorial neurodegenerative disease mainly caused by β-amyloid (Aβ) accumulation in the brain. Among the several factors that may concur to AD development, elevated cholesterol levels and brain cholesterol dyshomeostasis have been recognized to play a relevant role. Proprotein convertase subtilisin/kexin type 9 (PCSK9) is a protein primarily known to regulate plasma low-density lipoproteins (LDLs) rich in cholesterol and to be one of the main causes of familial hypercholesterolemia. In addition to that, PCSK9 is also recognized to carry out diverse important activities in the brain, including control of neuronal differentiation, apoptosis, and, importantly, LDL receptors functionality. Moreover, PCSK9 appeared to be directly involved in some of the principal processes responsible for AD development, such as inflammation, oxidative stress, and Aβ deposition. On these bases, PCSK9 management might represent a promising approach for AD treatment. The purpose of this review is to elucidate the role of PCSK9, whether or not cholesterol-related, in AD pathogenesis and to give an updated overview of the most innovative therapeutic strategies developed so far to counteract the pleiotropic activities of both humoral and brain PCSK9, focusing in particular on their potentiality for AD management. Full article
Show Figures

Figure 1

Figure 1
<p>PCSK9 involvement in brain cholesterol dyshomeostasis. Circulating PCSK9 prevents recycling of LDL receptors, inducing hypercholesterolemia, inflammation, and oxidative stress, thus leading to BBB damage that allows PCSK9 to enter into the brain. Inside the brain, PCSK9 affects receptors and transporters involved in astrocyte-synthetized cholesterol and cholesterol uptake by neurons. Abbreviations: ABC, ATP-binding cassette transporter; ApoE, Apolipoprotein E; ApoER2, Apolipoprotein E receptor 2; BBB, blood-brain barrier; CSF, cerebrospinal fluid; HDL, high-density lipoprotein; LDLR, low-density lipoprotein receptor; LRP1, lipoprotein receptor-related protein 1; PCSK9, Proprotein convertase subtilisin/kexin type 9; and VLDLR; very low-density lipoprotein receptor.</p>
Full article ">Figure 2
<p>Application of anti-PCSK9 pharmacological tools for AD treatment. Drugs developed to target circulating PCSK9 appear suitable for AD cure by counteracting amyloidogenesis and by reducing hypercholesterolemia, inflammation, and oxidative stress, thus preventing BBB damage. PCSK9-targeting drugs able to cross the BBB could be suggested to delay the neurodegenerative AD progression, exerting their activities directly inside the brain. Abbreviations: AAV, adeno-associated virus; EV, extracellular vesicle; mAb, monoclonal antibody; miRNA, microRNA; and siRNA, small interfering RNA.</p>
Full article ">Figure 3
<p>PCSK9 activities associated with Alzheimer’s disease onset and development and the relative factors involved. Abbreviations: ABCA1, ATP-binding cassette transporter A1; ApoE, Apolipoprotein E; ApoER2, Apolipoprotein E receptor 2; BACE1, beta-site amyloid precursor protein-cleaving enzyme-1; BBB, blood-brain barrier; LDLR, low-density lipoprotein receptor; LRP1, lipoprotein receptor-related protein 1; NF-κB, nuclear factor kappa B; NLRP3, NOD-like receptor protein 3; TLR4, Toll-like receptor 4; VLDLR; very low-density lipoprotein receptor.</p>
Full article ">
32 pages, 8535 KiB  
Review
Nanopipettes as a Potential Diagnostic Tool for Selective Nanopore Detection of Biomolecules
by Regina M. Kuanaeva, Alexander N. Vaneev, Petr V. Gorelkin and Alexander S. Erofeev
Biosensors 2024, 14(12), 627; https://doi.org/10.3390/bios14120627 - 19 Dec 2024
Viewed by 539
Abstract
Nanopipettes, as a class of solid-state nanopores, have evolved into universal tools in biomedicine for the detection of biomarkers and different biological analytes. Nanopipette-based methods combine high sensitivity, selectivity, single-molecule resolution, and multifunctionality. The features have significantly expanded interest in their applications for [...] Read more.
Nanopipettes, as a class of solid-state nanopores, have evolved into universal tools in biomedicine for the detection of biomarkers and different biological analytes. Nanopipette-based methods combine high sensitivity, selectivity, single-molecule resolution, and multifunctionality. The features have significantly expanded interest in their applications for the biomolecular detection, imaging, and molecular diagnostics of real samples. Moreover, the ease of manufacturing nanopipettes, coupled with their compatibility with fluorescence and electrochemical methods, makes them ideal for portable point-of-care diagnostic devices. This review summarized the latest progress in nanopipette-based nanopore technology for the detection of biomarkers, DNA, RNA, proteins, and peptides, in particular β-amyloid or α-synuclein, emphasizing the impact of technology on molecular diagnostics. By addressing key challenges in single-molecule detection and expanding applications in diverse biological areas, nanopipettes are poised to play a transformative role in the future of personalized medicine. Full article
(This article belongs to the Section Biosensors and Healthcare)
Show Figures

Figure 1

Figure 1
<p>Nanopore detection. (<b>a</b>) Experimental principle of nanopore detection. Two baths filled with an electrolyte solution, typically a buffered salt solution, are separated by a single nanopore (or nanopipette). Ag/AgCl electrodes are immersed in the bath and a constant voltage bias is applied across the nanopore. When a molecule translocates through the nanopore, the ionic current typically decreases. (<b>b</b>) Characteristics of translocation event. The amplitude and duration (dwell time) of the current reduction during translocation can provide critical insights about biomolecules. Created with Biorender.com.</p>
Full article ">Figure 2
<p>Comparison of biological-, solid-state-, and nanopipette-based nanopores: key materials, advantages, and limitations. Created with Biorender.com.</p>
Full article ">Figure 3
<p>Schematic representation of nanopipette fabrication and characterization. (<b>a</b>) Process of pulling a nanopipette. (<b>b</b>) Characteristics of the nanopipette, where <span class="html-italic">l</span> is the pore length, <span class="html-italic">d<sub>b</sub></span> and <span class="html-italic">d<sub>t</sub></span> are the base and tip diameters of the truncated cone, and α is the half-cone angle. Created with Biorender.com.</p>
Full article ">Figure 4
<p>Translocations of DNA origami. (<b>a</b>) Schematic representation of the concept of ECS. (<b>b</b>) ECS as a function of the DNA origami surface area for the four DNA nanostructure assembled. (<b>c</b>) ECS histograms of the DNA origami samples; from left to right: monomer sample, dimer sample, trimer sample, and 2 × 2 sample. Reproduced with permission from [<a href="#B90-biosensors-14-00627" class="html-bibr">90</a>]. Copyright 2022, Biophysical Society.</p>
Full article ">Figure 5
<p>Actin dynamics and its interaction with actin-binding drugs. (<b>a</b>) Schematic of experimental setup of protein detection using nanopipette-based nanopore. (<b>b</b>) (<b>Left</b>) protein models (actin with Latrunculin B bound, and actin with Swinholide A bound). (<b>Right</b>) typical current traces for actin bound to different filament inhibitors at 250 mV. (<b>c</b>) Scatterplots of current blockades vs. dwell times for both actin monomers and dimers with the same scale at 250 mV. Reproduced from [<a href="#B95-biosensors-14-00627" class="html-bibr">95</a>] with permission from the Royal Society of Chemistry.</p>
Full article ">Figure 6
<p>Nanopipette-based nanopore detection of glycoprotein. (<b>a</b>) Schematic illustration of the label-free monitoring of single-molecule glycoprotein−boronate affinity. Interaction via a 4-MPBA-modified nanopipette. (<b>b</b>) Current−time trace for the presence of 1 nM IgG in 100 mM KCl at +400 mV, (I) translocation of IgG molecules that do not interact with 4-MPBA, (II) translocation of IgG molecules interacted with 4-MPBA. (<b>c</b>) Two-dimensional contour plot of ΔI/I<sub>0</sub> vs. log(dwell time) of the single-IgG current blockade events. (<b>d</b>,<b>e</b>) Histograms showing the distributions of the logarithmic dwell time (<b>d</b>) and ΔI/I<sub>0</sub> (<b>e</b>). Red and blue indicate type I and type II signals, respectively. (<b>f</b>) Linear fit plot of the blockade event frequency (type II) and IgG concentrations from 0.02 to 5 nM. (<b>g</b>) Blockade event frequency (type II) of the selective detection of IgG in a blank solution, 1 nM IgG, and 1 nM IgG with different coexisting proteins 1 μM BSA, 1 μM SA, 1 μM LZ, 1 nM AFP, and 1 nM HRP. Reproduced with permission from [<a href="#B105-biosensors-14-00627" class="html-bibr">105</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 7
<p>Scheme of an aptamer-functionalized nexFET sensor. (<b>a</b>) Double-barrel quartz nanopipette: one barrel is hollow with a nanopore, while the other is filled with pyrolytic carbon. The pipette tip is coated with a thin carbon layer and features a core of PPy and a PPy–aptamer shell. (<b>b</b>) In the absence of a gate voltage, the core is positively charged, and the shell is slightly negatively charged, resulting in limited target protein binding. (<b>c</b>) Applying a positive gate voltage (V<sub>G</sub> = 400 mV) enhances the selective thrombin detection, increasing throughput and SNR. Reproduced from [<a href="#B108-biosensors-14-00627" class="html-bibr">108</a>]. Copyright 2020, the authors. Published by WILEY-VCH.</p>
Full article ">Figure 8
<p>Nanopore detection by a specific carrier. (<b>a</b>) Experimental principle of selective detection using molecular agent and a carrier; (<b>b</b>) example signal of selective detection. Created with Biorender.com.</p>
Full article ">Figure 9
<p>Nanopore detection of AChE using TDN. (<b>a</b>) Schematic illustration for the translocation of bare TDN-apt, AChE, and TDN-apt-AChE complex through the nanopipette and the corresponding signals. (<b>b</b>) Current−time traces of the translocation of 150 fM TDN-apt interacting with 100 fM AChE and 100 fM of other interferent proteins, as well as their mixture. Reprinted with permission from [<a href="#B110-biosensors-14-00627" class="html-bibr">110</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 10
<p>Nanopore detection of lysozyme by AuNPs. (<b>a</b>) Structure of AuNPs functionalized with an LBA (1) for detection of lysozyme (2). (<b>b</b>) Current–time traces of the AuNP-LBA-lysozyme complex at 600 mV with 100 mM KCl. Reproduced from [<a href="#B112-biosensors-14-00627" class="html-bibr">112</a>]. Copyright 2017, Royal Society of Chemistry.</p>
Full article ">Figure 11
<p>Detection of miRNA-141 and PCT by AuNPs. (<b>a</b>) Schematic representation of AuNP monomer miR-141-3p molecular probes with representative individual events (scale bar: vertical 50 pA, horizontal 20 µs), along with the associated statistics. (<b>b</b>) Conjugated dimers with miRNA-141 linked between 2 NP monomers. (<b>c</b>) AuNP monomeric antibody molecular probes with individual translocation events (scale bar: vertical 50 pA, horizontal 20 µs), along with associated statistics. (<b>d</b>) Conjugated antibody dimers with PCT (an antigen). Reproduced from [<a href="#B113-biosensors-14-00627" class="html-bibr">113</a>]. Copyright 2021, John Wiley and Sons.</p>
Full article ">Figure 12
<p>Double-barrel nanopipette-based system for the detection of dopamine, serotonin, and K<sup>+</sup>. (<b>a</b>) Schematic illustration of the measurement principle based on a dual nanopipette. Ionic current recordings of the translocations of 20 nm Au-PEG NPs in 100 mM KCl at 400 mV performed in a single nanopipette (<b>b</b>) and a dual nanopipette (<b>c</b>) (stars highlight representative events in dashed boxes). Reprinted with permission from [<a href="#B115-biosensors-14-00627" class="html-bibr">115</a>]. Copyright 2022, American Chemical Society.</p>
Full article ">Figure 13
<p>Nanopore electro-optical approach to detect thrombin. (<b>a</b>) Schematic of the electro-optical configuration, where a nanopore is integrated with a single-molecule fluorescence confocal microscope. (<b>b</b>) MBs are hybridized to a DNA carrier for the single-molecule detection of small oligonucleotides or proteins. Top: without target binding, the signal is only observed in the electrical detection channel. Bottom: when bound to a complementary DNA or protein, a synchronized signal is observed in both channels due to the opening of the MB and the increased distance between the fluorophore and the quencher probes. (<b>c</b>) Photon and current time traces are shown for the translocation of (i) thrombin in 5% serum, (ii) MB–carrier in 5% serum, and (iii) MB–carrier bound to thrombin in 5% serum. (iv) Percent synchronization between the optical and electrical channels for thrombin bound to the MB–carrier. Reprinted from [<a href="#B116-biosensors-14-00627" class="html-bibr">116</a>]. Copyright 2019, the authors.</p>
Full article ">Figure 14
<p>Multiplexed detection of SARS-CoV-2 viral proteins and RNA fragments. (<b>a</b>) Schematic of multiplexed detection of SARS-CoV-2 viral proteins and RNA fragments. (<b>b</b>) Ion current–time traces for 10 kbp dsDNA probes encoded with an S protein aptamer without bound S protein (i), and when the S protein is bound, a secondary peak occurs at the end (ii). (<b>c</b>) Similarly, 10 kbp dsDNA probes encoded with an N protein aptamer (without bound with N protein (i)), display a secondary peak in the middle upon N protein binding (ii). (<b>d</b>) A 9.1 kbp DNA probe encoded with sequences complementary to the ORF1b, S, and N genes (i) enables RNA fragment detection, with specific secondary peaks corresponding to each gene (ii). Reprinted from [<a href="#B41-biosensors-14-00627" class="html-bibr">41</a>]. Copyright 2019, the authors.</p>
Full article ">Figure 15
<p>Mapping of recognition sites on DNA. (<b>a</b>) Single dCas9 probes bound to 3.6 kbp DNA translocating through a nanopore. (<b>b</b>) Double and triple dCas9 probe barcodes on full-length λ-DNA. (i) Example events with two and three peaks due to binding of the double-probe barcode and triple-probe barcode, respectively. (ii) Raw data comparing the number of peaks counted per event after the addition of a single probe, double probe, and triple probe. Reprinted with permission from [<a href="#B117-biosensors-14-00627" class="html-bibr">117</a>]. Copyright 2019, American Chemical Society.</p>
Full article ">Figure 16
<p>Single-molecule monitoring of Aβ<sub>1–42</sub> monomers. (<b>a</b>) Schematic representation of the single-molecule monitoring of Aβ<sub>1–42</sub> monomers: (i) event scatter plots and (ii) histograms of the current amplitude versus dwell time for the translocation of Aβ<sub>1–42</sub> monomers. (<b>b</b>) Single-molecule monitoring of Aβ<sub>1–42</sub> oligomers: (i) event scatter plots and (ii) histograms of the current amplitude versus dwell time for the translocation of Aβ<sub>1–42</sub> oligomers. (<b>c</b>) Single-molecule monitoring of Aβ<sub>1–42</sub> fibers: (i) event scatter plots and (ii) histograms of the current amplitude versus dwell time for the translocation of Aβ<sub>1–42</sub> fibers. Reproduced from [<a href="#B127-biosensors-14-00627" class="html-bibr">127</a>] with permission from the Royal Society of Chemistry.</p>
Full article ">Figure 17
<p>Concept of real-time fast amyloid seeding and translocation (RT-FAST). (<b>a</b>) Scheme showing the two parts of a nanopipette: the reservoir where the αS seeds are amplified and the sensor where αS seeds are detected. (<b>b</b>) Illustration of the RT-FAST experiments. (<b>c</b>) Example of a current trace record extracted from different nanopipettes for reference (light blue), the control not seeded (blue), and the sample seeded with αS WT (red) and A53T mutant (yellow). Reproduced from [<a href="#B132-biosensors-14-00627" class="html-bibr">132</a>]. Copyright 2022, the authors.</p>
Full article ">
27 pages, 2940 KiB  
Review
Understanding the Molecular Impact of Physical Exercise on Alzheimer’s Disease
by Alba Cantón-Suárez, Leticia Sánchez-Valdeón, Laura Bello-Corral, María J. Cuevas and Brisamar Estébanez
Int. J. Mol. Sci. 2024, 25(24), 13576; https://doi.org/10.3390/ijms252413576 - 18 Dec 2024
Viewed by 500
Abstract
Alzheimer’s disease is one of the most common neurodegenerative diseases, characterized by a wide range of neurological symptoms that begin with personality changes and psychiatric symptoms, progress to mild cognitive impairment, and eventually lead to dementia. Physical exercise is part of the non-pharmacological [...] Read more.
Alzheimer’s disease is one of the most common neurodegenerative diseases, characterized by a wide range of neurological symptoms that begin with personality changes and psychiatric symptoms, progress to mild cognitive impairment, and eventually lead to dementia. Physical exercise is part of the non-pharmacological treatments used in Alzheimer’s disease, as it has been shown to delay the neurodegenerative process by improving the redox state in brain tissue, providing anti-inflammatory effects or stimulating the release of the brain-derived neurotrophic factor that enhances the brain structure and cognitive performance. Here, we reviewed the results obtained from studies conducted in both animal models and human subjects to comprehend how physical exercise interventions can exert changes in the molecular mechanisms underlying the pathophysiological processes in Alzheimer’s disease: amyloid β-peptide pathology, tau pathology, neuroglial changes, mitochondrial dysfunction, and oxidative stress. Physical exercise seems to have a protective effect against Alzheimer’s disease, since it has been shown to induce positive changes in some of the biomarkers related to the pathophysiological processes of the disease. However, additional studies in humans are necessary to address the current lack of conclusive evidence. Full article
(This article belongs to the Special Issue Molecular Insights into the Role of Exercise in Disease and Health)
Show Figures

Figure 1

Figure 1
<p>Non-modifiable and modifiable risk factors for AD. Age, gender, and genetics are considered non-modifiable factors in AD. Conversely, psychosocial factors (cognitive activity, educational attainment, depression, stress, or sleep disturbances), environmental factors (smoking, alcohol intake, diet, or physical exercise), and pre-existing diseases (diabetes, hypertension, obesity, or cardiovascular diseases) are considered modifiable factors in the development of AD.</p>
Full article ">Figure 2
<p>AD main pathophysiological mechanisms. The presence of amyloid plaques and NFTs in the brain of AD patients alongside changes in neuroglial cells, increased levels of oxidative stress, and mitochondrial damage and dysfunction are the main pathophysiological mechanisms present in AD.</p>
Full article ">Figure 3
<p>APP processing pathways. APP is cleaved by α-secretase and β- and γ-secretases into soluble fragments sAPPα and sAPPβ, respectively. While sAPPα does not lead to plaque formation (non-amyloidogenic pathway), Aβ peptides aggregate to form amyloid plaques (amyloidogenic pathway).</p>
Full article ">Figure 4
<p>Tau pathology. Hyperphosphorylated tau cannot attach properly to microtubules, leading to microtubule disintegration and the aggregation of tau protein into paired helical filaments and NFTs.</p>
Full article ">Figure 5
<p>Microglia phenotypes. Depending on the stimuli received, resting microglia can change into activated phases with two different phenotypes: M2 or anti-inflammatory phenotype and M1 or pro-inflammatory phenotype.</p>
Full article ">Figure 6
<p>Mitochondria fusion and fission processes.</p>
Full article ">Figure 7
<p>Schematic representation of the molecular mechanisms modified by PE in AD. Data extracted from the studies presented in <a href="#ijms-25-13576-t001" class="html-table">Table 1</a> and <a href="#ijms-25-13576-t002" class="html-table">Table 2</a>.</p>
Full article ">
28 pages, 2534 KiB  
Review
Irisin: A Multifaceted Hormone Bridging Exercise and Disease Pathophysiology
by Ilaria Paoletti and Roberto Coccurello
Int. J. Mol. Sci. 2024, 25(24), 13480; https://doi.org/10.3390/ijms252413480 - 16 Dec 2024
Viewed by 501
Abstract
The fibronectin domain-containing protein 5 (FNDC5), or irisin, is an adipo-myokine hormone produced during exercise, which shows therapeutic potential for conditions like metabolic disorders, osteoporosis, sarcopenia, obesity, type 2 diabetes, and neurodegenerative diseases, including Alzheimer’s disease (AD). This review explores its potential across [...] Read more.
The fibronectin domain-containing protein 5 (FNDC5), or irisin, is an adipo-myokine hormone produced during exercise, which shows therapeutic potential for conditions like metabolic disorders, osteoporosis, sarcopenia, obesity, type 2 diabetes, and neurodegenerative diseases, including Alzheimer’s disease (AD). This review explores its potential across various pathophysiological processes that are often considered independent. Elevated in healthy states but reduced in diseases, irisin improves muscle–adipose communication, insulin sensitivity, and metabolic balance by enhancing mitochondrial function and reducing oxidative stress. It promotes osteogenesis and mitigates bone loss in osteoporosis and sarcopenia. Irisin exhibits anti-inflammatory effects by inhibiting NF-κB signaling and countering insulin resistance. In the brain, it reduces amyloid-β toxicity, inflammation, and oxidative stress, enhancing brain-derived neurotrophic factor (BDNF) signaling, which improves cognition and synaptic health in AD models. It also regulates dopamine pathways, potentially alleviating neuropsychiatric symptoms like depression and apathy. By linking physical activity to systemic health, irisin emphasizes its role in the muscle–bone–brain axis. Its multifaceted benefits highlight its potential as a therapeutic target for AD and related disorders, with applications in prevention, in treatment, and as a complement to exercise strategies. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>Irisin has pleiotropic effects, as it exerts multiple, diverse biological actions in different tissues and organ systems, meaning its effects are not limited to one specific function but span several physiological processes that impact metabolism, muscle function, fat storage, bone health, and brain activity. In particular, in skeletal muscle, irisin triggers the (MAPK)-PGC-1α pathway, thus improving oxidative phosphorylation and consequently mitochondrial respiration, together with an increase in GLUT4 translocation. At the same time, in cardiac tissue, GLUT4 translocation appears downstream of the activation of the PI3K/PKB/Akt pathway, with the final result of improving glucose uptake. In relationship to the liver, downstream, the PI3K trigger Akt/GSK3 is then activated, leading to gluconeogenesis reduction and the concomitant induction of glycogen synthesis. Altogether, with the stimulation of all the aforementioned pathways, irisin can promote sensitization to glucose in insulin-dependent tissue, like the heart, skeletal muscle, and the liver (<b>pink lines</b>). This adipo-myokine plays a crucial role in the hepatic metabolism, as AMPK phosphorylation exerts anti-inflammatory potential by decreasing NLRP3 inflammasome and NFKβ (<b>green lines</b>). Through AMPK phosphorylation, irisin also participates in autophagic-related mechanisms via mTOR inhibition promoting osteoblastogenesis by the parallel activation of Wtn/β catenin signaling. Furthermore, irisin is involved in osteogenesis involving ERK signaling via p38 MAPK, the same pathway mediating GLUT4 translocation in muscle tissue (<b>blue lines</b>). Irisin plays a crucial role not only in muscle–bone crosstalk but also acts on many other peripheral organs. Moreover, irisin shows antioxidant effects in the intestine and pancreas, as it can upregulate UCP2, reducing both oxidative stress and ER stress in intestinal cells, additionally increasing mitochondrial function and biogenesis (<b>violet lines</b>). Interestingly, irisin exerts a protective action on pancreatic β-cells to stimulate insulin synthesis and glucose-induced insulin secretion. Specifically, irisin can block the apoptotic effects induced by selective saturated fatty acids in β-cells via the Akt and Bcl-2 signaling pathway (<b>red lines</b>). Lastly, irisin acts as antihypertensive hormone, since it reduces blood pressure by activating the hypothalamic factor Nrf2, having a more selective effect on PVN nuclei for its antioxidant and anti-inflammatory action. Thus, this hormone cannot be simply identified for its single effects on specific organs, since its beneficial activity is provided by the synergistic combination of its action on multiple organs and systems.</p>
Full article ">Figure 2
<p>Skeletal muscle and bone are dynamic, adaptable tissues that function as endocrine and paracrine organs, maintaining tight reciprocal control and communication. Osteoporosis and sarcopenia often co-exist in obese patients with visceral adiposity, a condition known as osteosarcopenic obesity. This figure illustrates how irisin helps counteract and rebalance these pathological conditions by promoting muscle growth and bone formation. Physical exercise increases circulating irisin levels, as well as PGC-1α and FNDC5, leading to beneficial effects on bone health, particularly by stimulating pro-osteoblastic mechanisms and enhancing bone formation. Adiposity, characterized as a state of chronic low-grade inflammation, involves excessive lipid infiltration in skeletal muscle, which induces lipotoxicity and inflammation associated with myosteatosis. This fat deposition triggers the release of pro-inflammatory cytokines, affecting the osteoprotegerin pathway and promoting osteoclastogenesis and bone resorption. Irisin, by reducing inflammation and lipotoxicity (indicated by green arrows), can decrease bone resorptive factors, activate the Wnt/β-catenin signaling pathway, and stimulate osteoblast differentiation. Lipid accumulation also contributes to insulin resistance. Irisin reduces adipose tissue inflammation, improving the inflammatory impact of lipid deposits and lowering the risk of insulin resistance. Since insulin signaling deficiency is linked to alterations in bone microarchitecture, β-cell dysfunction is associated with osteoporosis, highlighting the direct relationship between insulin resistance and bone metabolism. Irisin acts as a hormonal messenger capable of counteracting insulin resistance and myosteatosis, thereby connecting insulin metabolism with muscle and bone homeostasis. Specifically, irisin can decrease NLRP3 inflammasome activity, promote GLUT4 translocation in muscle, and upregulate GLUT4 via the AMPK signaling pathway.</p>
Full article ">Figure 3
<p>The figure depicts the capacity of irisin to affect multiple pathways involved in age-associated neurodegeneration and mechanisms underlying AD pathogenesis. Irisin serves as a crucial link between muscle contraction, myokine secretion, and brain function, embodying the concept of the muscle–brain axis. It has been shown to have potent (1) anti-inflammatory and (2) neuroprotective effects, such as protecting against Aβ neurotoxicity, reducing the release of pro-inflammatory cytokines, like IL-1β and IL-6, and inhibiting the expression of COX-2 and NF-κB in astrocytes. Irisin’s neuroprotective actions extend to preventing neuronal damage in conditions like middle cerebral artery occlusion and brain infarction, as well as mitigating microglial activation, neutrophil infiltration, and the expression of TNF-α and IL-6 via the ERK1/2 and Akt signaling pathways. Additionally, it counters the TLR4/MyD88-mediated neuroinflammatory response. The integrin αV/β5 receptor, which is highly expressed in microglia, plays a key role in irisin’s effects. Irisin has been shown to promote the shift from an M1 pro-inflammatory phenotype to an M2-like anti-inflammatory phenotype in microglia, leading to changes in their morphology. This shift is associated with an increase in AMPK phosphorylation and the expression of the anti-apoptotic protein Bcl-2. The interaction between irisin and BDNF signaling is also essential in neurodegenerative diseases, particularly Alzheimer’s disease (AD). For example, FNDC5 overexpression can mitigate the inhibitory effects of Aβ<sub>1-42</sub> oligomers on BDNF expression. Exercise has been linked to increased (3) lactate levels and elevated hippocampal BDNF expression, with SIRT1 histone deacetylase activating the PGC-1α/FNDC5 signaling pathway to induce BDNF expression. Irisin’s (3) antidepressant potential is particularly relevant for managing neuropsychiatric symptoms in AD. Its ability to enhance dopaminergic activity (i.e., right side of the figure) may help address the pathophysiology of depression. Specifically, irisin-induced BDNF expression can activate dopamine D3 receptor-mediated signaling through the Akt and ERK pathways, fine-tuning dopamine release and contributing to a reduction in Aβ deposits and the overall Aβ burden.</p>
Full article ">
24 pages, 3290 KiB  
Review
Targeting Iron Responsive Elements (IREs) of APP mRNA into Novel Therapeutics to Control the Translation of Amyloid-β Precursor Protein in Alzheimer’s Disease
by Mateen A. Khan
Pharmaceuticals 2024, 17(12), 1669; https://doi.org/10.3390/ph17121669 - 11 Dec 2024
Viewed by 618
Abstract
The hallmark of Alzheimer’s disease (AD) is the buildup of amyloid-β (Aβ), which is produced when the amyloid precursor protein (APP) misfolds and deposits as neurotoxic plaques in the brain. A functional iron responsive element (IRE) RNA stem loop is encoded by the [...] Read more.
The hallmark of Alzheimer’s disease (AD) is the buildup of amyloid-β (Aβ), which is produced when the amyloid precursor protein (APP) misfolds and deposits as neurotoxic plaques in the brain. A functional iron responsive element (IRE) RNA stem loop is encoded by the APP 5′-UTR and may be a target for regulating the production of Alzheimer’s amyloid precursor protein. Since modifying Aβ protein expression can give anti-amyloid efficacy and protective brain iron balance, targeted regulation of amyloid protein synthesis through modulation of 5′-UTR sequence function is a novel method for the prospective therapy of Alzheimer’s disease. Numerous mRNA interference strategies target the 2D RNA structure, even though messenger RNAs like tRNAs and rRNAs can fold into complex, three-dimensional structures, adding even another level of complexity. The IRE family is among the few known 3D mRNA regulatory elements. This review seeks to describe the structural and functional aspects of IREs in transcripts, including that of the amyloid precursor protein, that are relevant to neurodegenerative diseases, including AD. The mRNAs encoding the proteins involved in iron metabolism are controlled by this family of similar base sequences. Like ferritin IRE RNA in their 5′-UTR, iron controls the production of APP in their 5′-UTR. Iron misregulation by iron regulatory proteins (IRPs) can also be investigated and contrasted using measurements of the expression levels of tau production, Aβ, and APP. The development of AD is aided by iron binding to Aβ, which promotes Aβ aggregation. The development of small chemical therapeutics to control IRE-modulated expression of APP is increasingly thought to target messenger RNAs. Thus, IRE-modulated APP expression in AD has important therapeutic implications by targeting mRNA structures. Full article
Show Figures

Figure 1

Figure 1
<p>Proposed model for iron-induced translation modulation by the IRP/IRE signaling pathway. Low cellular iron promotes the binding of IRP, release of eIF4F and blocking ribosome binding to IRE in the UTRs of APP thus preventing APP translation. In cellular iron overload, iron displaces IRPs from the IRE mRNA. IRP1 dissociates, allowing ribosome and initiation factor eIF4F binding and translation of APP proceeds. Black arrows showing either eIF/IRP interacting or dissociating from IRE stem-loop structure of RNA. Red cross means blocking of ribosome binding.</p>
Full article ">Figure 2
<p>Alzheimer’s disease with iron dysregulation in the brain. How the production of APP and the promotion of amyloid plaques in AD are caused by an increase in the labile iron pool (LIP) in neurons.</p>
Full article ">Figure 3
<p>Amyloid β configuration and accumulation. (<b>A</b>) The Aβ<sub>42</sub> primary amino acid sequence. (<b>B</b>) Amyloid-β peptide structures interact with Aβ aggregates during the elongation process (modified from Ref. [<a href="#B97-pharmaceuticals-17-01669" class="html-bibr">97</a>]). (<b>C</b>) The aggregation of monomers of Aβ into higher-order oligomers, fibrils, protofibrils, and amyloid plaques is triggered by iron.</p>
Full article ">Figure 4
<p>APP IRE RNA 5′-UTRs were predicted to fold into stable RNA stem loops like the 5′-UTR-specific IRE in ferritin-H and ferritin-L transcript. (<b>A</b>) APP, ferritin-H, and ferritin-L stem loop IRE structures. (<b>B</b>) After comparing sequences encoding the 5′-UTR-specific IRE stem loop with ferritin IRE (which were displayed in two clusters of greater than 70% sequence similarity), the APP IRE was discovered. Key IRE motifs are highlighted in yellow. (<b>C</b>) APP, ferritin-H, and ferritin-L transcript maps of the 5′-UTR IRE stem loops.</p>
Full article ">Figure 5
<p>APP iron responsive element RNA structural models. Conserved APP IRE (<b>A</b>) secondary and (<b>B</b>) tertiary structure. (<b>C</b>) APP IRE RNA binding to IRP1. APP IRE RNA: IRP1 complex showing bulge bases and triloop bases flipped out of the helix and making deep contacts in IRP1 protein pockets (modification of figure originally published in ref. [<a href="#B40-pharmaceuticals-17-01669" class="html-bibr">40</a>]).</p>
Full article ">Figure 6
<p>Iron selectively weakens APP IRE RNA/IRP1 interactions. IRP1 binds to the APP IRE RNA with nM affinity in the absence and of iron. Binding curve (protein fluorescence quenching) and conserved IRE secondary structure are prepared from binding data originally published in ref. [<a href="#B40-pharmaceuticals-17-01669" class="html-bibr">40</a>].</p>
Full article ">Figure 7
<p>Diagrammatic representation of the translation of the amyloid-β therapeutic target APP IRE RNA. (<b>A</b>) IRP can bind to the IRE loop in the 5′-UTR of the APP mRNA and stop amyloid translation when there is a low iron content. (<b>B</b>) In addition to allowing eIF4F and ribosome binding to APP mRNA, which overexpresses neurotoxic amyloid precursor protein, high iron levels also dissociate IRP. (<b>C</b>) Small-molecule RNA or peptide that binds to APP IRE RNA and inhibits the synthesis of amyloid proteins, hence inhibiting eIF4F and ribosomes.</p>
Full article ">
16 pages, 2083 KiB  
Review
Anthranilic Acid–G-Protein Coupled Receptor109A–Cytosolic Phospholipase A2–Myelin–Cognition Cascade: A New Target for the Treatment/Prevention of Cognitive Impairment in Schizophrenia, Dementia, and Aging
by Gregory Oxenkrug
Int. J. Mol. Sci. 2024, 25(24), 13269; https://doi.org/10.3390/ijms252413269 - 10 Dec 2024
Viewed by 575
Abstract
Cognitive impairment is a core feature of neurodevelopmental (schizophrenia) and aging-associated (mild cognitive impairment and Alzheimer’s dementia) neurodegenerative diseases. Limited efficacy of current pharmacological treatments warrants further search for new targets for nootropic interventions. The breakdown of myelin, a phospholipids axonal sheath that [...] Read more.
Cognitive impairment is a core feature of neurodevelopmental (schizophrenia) and aging-associated (mild cognitive impairment and Alzheimer’s dementia) neurodegenerative diseases. Limited efficacy of current pharmacological treatments warrants further search for new targets for nootropic interventions. The breakdown of myelin, a phospholipids axonal sheath that protects the conduction of nerve impulse between neurons, was proposed as a neuropathological abnormality that precedes and promotes the deposition of amyloid-β in neuritic plaques. The present review of the recent literature and our own pre- and clinical data suggest (for the first time) that the anthranilic acid (AA)-induced activation of microglial-expressed G-protein coupled receptor (GPR109A) inhibits cytosolic phospholipase A2 (cPLA2), an enzyme that triggers the degradation of myelin and consequently attenuates cognitive impairment. The present review suggests that the up-regulation of AA formation is a sex-specific compensatory (adaptive) reaction aimed to prevent/treat cognitive impairment. The AA–GPR109A–cPLA2–myelin–cognition cascade suggests new nootropic interventions, e.g., the administration of pegylated kynureninase, an enzyme that catalyzes AA formation from Kynurenine (Kyn), a tryptophane catabolite; pegylated interferon-alpha; central and peripheral Kyn aminotransferase inhibitors that increase availability of Kyn as a substrate for AA formation; and vagus nerve stimulation. The cascade predicts nootropic activity of exogenous GPR109A agonists that were designed and underwent clinical trials (unsuccessful) as anti-dyslipidemia agents. The proposed cascade might contribute to the pathogenesis of cognitive impairment. Data on AA in neurodegenerative disorders are scarce, and the proposed cascade needs further exploration in pre- and clinical studies Full article
Show Figures

Figure 1

Figure 1
<p>Proposed GPR109A downregulation in cognitive impairment. Abbreviations: GPR109A: G-protein coupled receptor; cPLA2: cytosolic phospholipase A2. <span class="html-fig-inline" id="ijms-25-13269-i001"><img alt="Ijms 25 13269 i001" src="/ijms/ijms-25-13269/article_deploy/html/images/ijms-25-13269-i001.png"/></span> up-regulation <span class="html-fig-inline" id="ijms-25-13269-i002"><img alt="Ijms 25 13269 i002" src="/ijms/ijms-25-13269/article_deploy/html/images/ijms-25-13269-i002.png"/></span> down-regulation.</p>
Full article ">Figure 2
<p>Tryptophan–kynurenine–anthranilic acid pathway in humans. Abbreviations: Trp: tryptophan; Kyn: kynurenine; 3HK: 3-hydroxykynurenine; AA: anthranilic acid; KYNA: kynurenic acid; QUIN: quinolinic acid; NAD<sup>+</sup>: nicotinamide adenine dinucleotide; KMO: kynurenine 3-monooxygenase; KYNU: kynureninase; KAT: kynurenine aminotransferase; GPR109A: G-protein coupled receptor; NMDAR: N-methyl-D-aspartate receptor. * denotes Michaelis–Menten constant.</p>
Full article ">Figure 3
<p>Proposed mechanism of nootropic effect of anthranilic acid. Abbreviations: AA: anthranilic acid; GPR109A: G-protein coupled receptor; cPLA2: cytosolic phospholipase 2. <span class="html-fig-inline" id="ijms-25-13269-i003"><img alt="Ijms 25 13269 i003" src="/ijms/ijms-25-13269/article_deploy/html/images/ijms-25-13269-i003.png"/></span> Up-regulation <span class="html-fig-inline" id="ijms-25-13269-i004"><img alt="Ijms 25 13269 i004" src="/ijms/ijms-25-13269/article_deploy/html/images/ijms-25-13269-i004.png"/></span> Down-regulation.</p>
Full article ">Figure 4
<p>Chemical structures of anthranilic acid and benzoate sodium. Lennerz et al. [<a href="#B75-ijms-25-13269" class="html-bibr">75</a>]; Subramanian et al. [<a href="#B76-ijms-25-13269" class="html-bibr">76</a>].</p>
Full article ">Figure 5
<p>Kynurenine downstream catabolism in schizophrenia (<b>A</b>) and under KAT inhibition in schizophrenia (<b>B</b>). Abbreviations: Trp: tryptophan; Kyn: kynurenine; KYNA: kynurenic acid; 3HK: 3-hydroxykynurenine; AA: anthranilic acid; NAD+: nicotinamide adenine dinucleotide; GPR109A: G-protein coupled receptor 109A; NMDAR: N-methyl-D-aspartate receptor.</p>
Full article ">
Back to TopTop