Next Article in Journal
From Differential DNA Methylation in COPD to Mitochondria: Regulation of AHRR Expression Affects Airway Epithelial Response to Cigarette Smoke
Next Article in Special Issue
Cathelicidin-Related Antimicrobial Peptide Negatively Regulates Bacterial Endotoxin-Induced Glial Activation
Previous Article in Journal
Changes in Gut Microbiota and Systemic Inflammation after Synbiotic Supplementation in Patients with Systemic Lupus Erythematosus: A Randomized, Double-Blind, Placebo-Controlled Trial
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Hidden Role of Non-Canonical Amyloid β Isoforms in Alzheimer’s Disease

1
Department of Informatics and Microsystems Technology, University of Applied Sciences Kaiserslautern, D-66482 Zweibruecken, Germany
2
Department of Neurogenetics, Max Planck Institute for Multidisciplinary Sciences, D-37075 Goettingen, Germany
3
Department of Psychiatry and Psychotherapy, University Medical Centre of the Johannes Gutenberg University, D-55131 Mainz, Germany
*
Author to whom correspondence should be addressed.
Cells 2022, 11(21), 3421; https://doi.org/10.3390/cells11213421
Submission received: 23 August 2022 / Revised: 18 October 2022 / Accepted: 20 October 2022 / Published: 29 October 2022

Abstract

:
Recent advances have placed the pro-inflammatory activity of amyloid β (Aβ) on microglia cells as the focus of research on Alzheimer’s Disease (AD). Researchers are confronted with an astonishing spectrum of over 100 different Aβ variants with variable length and chemical modifications. With the exception of Aβ1-42 and Aβ1-40, the biological significance of most peptides for AD is as yet insufficiently understood. We therefore aim to provide a comprehensive overview of the contributions of these neglected Aβ variants to microglia activation. First, the impact of Aβ receptors, signaling cascades, scavenger mechanisms, and genetic variations on the physiological responses towards various Aβ species is described. Furthermore, we discuss the importance of different types of amyloid precursor protein processing for the generation of these Aβ variants in microglia, astrocytes, oligodendrocytes, and neurons, and highlight how alterations in secondary structures and oligomerization affect Aβ neurotoxicity. In sum, the data indicate that gene polymorphisms in Aβ-driven signaling pathways in combination with the production and activity of different Aβ variants might be crucial factors for the initiation and progression of different forms of AD. A deeper assessment of their interplay with glial cells may pave the way towards novel therapeutic strategies for individualized medicine.

1. Introduction

The neuroinflammatory activity of microglia has, since its discovery, been suspected to contribute to Alzheimer’s Disease (AD) [1,2,3]. Early reports referred to an invasion of reactive microglia at the site of senile plaques [4,5]. Moreover, it was observed that both microglia and astrocytes display exacerbated pro-inflammatory activity in disease-affected brain areas [6]. This led to the theory that glial cells are either directly involved in the pathogenesis of AD or are the immune system’s defensive shield against it [1]. Recent advances have solidified both presumptions and thus placed the inflammatory activity of glial cells as the focus of AD research [3].
AD is closely linked to the deposition of misfolded amyloid peptides in plaques and the occurrence of neurofibrillary tangles consisting of hyperphosphorylated tau protein [7]. While these discoveries were made more than three decades ago [8], the precise role of Aβ in neuroinflammation and neurodegeneration is still incompletely understood [9]. Progress is hampered by some enormous challenges in Aβ research, ranging from a sophisticated peptide process involving multiple proteases [10,11], to complex physiochemical properties of the resulting peptides [12], leading to different oligomeric structures [13] that ultimately elicit miscellaneous physiological effects on neurons, glia, and immune cells [3,7]. We now know that glial cells express a wide array of receptors such as triggering receptor expressed on myeloid cells 2 (TREM2), toll-like receptors (TLRs), and formyl peptide receptors (FPRs), which are capable of directly interacting with extracellular Aβ (see Figure 1). Several other molecules, such as NOD-, LRR- and pyrin domain-containing protein 3 (NLRP3), transient receptor potential cation channel, subfamily M, member 2 (TRPM2), and cluster of differentiation 36 (CD36) are further key modulators of the inflammatory Aβ cascades [14]. More than 100 Aβ peptides with varying lengths, structure, and chemical modifications have been described [10,11,15]. However, most studies focused only on the two most common variants, Aβ1-42 and Aβ1-40. Thus, the effects of many other fragments that are generated under physiological conditions and are present in significant amounts in affected brain areas remain to be elucidated. We here aim to provide an overview of the wide range of interactions between glial cells and these neglected Aβ variants in the inflammatory cycle of AD. In this context, we discuss the influence of different glia cell types and their cell-surface receptors on APP processing, extracellular cleavage, and chemical modifications of different Aβ species. Finally, we discuss the impact of these variants on neuroinflammation and summarize the current knowledge of gene variants in Aβ receptors on the perception of different Aβ species.

2. The Role of Glia in Neuroinflammation

Today we know that an aberrant behavior of both microglia and astrocytes occurs even at early stages of AD where typical cognitive symptoms and neuronal decay are not yet observed [19,20,21].
Microglia are the driving force behind the immune defense of the central nervous system (CNS) [3]. As highly motile cells, they survey their environment for pathological stimuli—either from exogenous sources such as bacteria and viruses, or from endogenous structures such as debris of damaged cells and misfolded proteins [22]. Due to its complex structure, the CNS and especially its neuronal network is highly vulnerable to pro-inflammatory stress and cell damage [23]. Therefore, the inflammatory response of microglia has to strike a finely tuned balance between an effective removal of the pathogens and the remaining viability of the surrounding neuronal cells [1,23]. This requires a strictly regulated network of cell-surface receptors and intracellular signaling molecules that carefully control secretion of pro- and anti-inflammatory cytokines and chemokines, generation of oxidative stress through the release of reactive oxygen species (ROS), and phagocytic uptake and degradation of pathogens and cell debris [24]. In addition to these functions, microglia also contribute to the maintenance of healthy brain homeostasis [24,25]. Here, they support neuronal survival and proliferation through constant cross talk with neighboring neurons and astrocytes, and release of trophic factors such as insulin growth factor 1 (IGF-1) or transforming growth factor β (TGFβ) [26,27]. Microglia also help to coordinate the functional state of neurons [28] because they participate in the development of neuronal networks by removing dysfunctional synapses through synaptic pruning [29] and by promoting formation of new synapses through the release of factors such as brain-derived neurotrophic factor (BDNF) [30,31]. Because of their multifaceted roles, microglia have an enormous capability to rapidly change their expression profile and phenotype depending on the encountered stimuli [32]. In a healthy environment, they possess a ramified morphology with wide soma and long branches that help to sense their local environment in order to pursue supportive and vigilant roles. After detection of harmful stimuli, they undergo significant molecular and morphological changes to assume a reactive phenotype with an amoeboid shape and initialize pro-inflammatory processes that help to eliminate pathogens from the CNS [3]. Historically, the amoeboid phenotype has been classified as the pro-inflammatory M1-type and the ramified morphology as the anti-inflammatory M2-type [3]. Recent studies demonstrated that these classifications are an oversimplification because microglia can adopt several different pro- and anti-inflammatory states with distinct gene expression profiles that depend on their local microenvironment and ageing of individual cells [32,33]. In addition, RNA-seq analysis of transgenic AD mice provided evidence for the existence of multiple Disease-Associated Microglia (DAM) subsets, showing unique expression profiles [34] in different forms of neuroinflammation [34,35]. For the sake of clarity, we still refer to the sum of all pro-inflammatory microglial cell types as “reactive microglia”.
Clear evidence indicates that reactive microglia can detect and respond to Aβ. Over the last decade it has become apparent that there is no single interaction partner for extracellular Aβ. Instead, reactive microglia display a large number of different receptors that are potential binding partners. Depending on the environment, disease type, and age of the respective human host, different subsets of reactive microglia exist that express discrete combinations of these receptors that modulate their inflammatory responses. The currently known direct Aβ interaction partners expressed in microglia include structurally diverse molecules such as triggering receptor expressed on myeloid cells 2 (TREM2), its co-receptor sialic acid binding Ig-like lectin 3 (CD33), cluster of differentiation 36 (CD36), toll-like receptors (TLRs), formyl peptide receptors (FPRs), receptor for advanced glycation endproducts (RAGE), chemokine-like receptor 1 (CMKLR1), macrophage receptor with collagenous structure (MARCO), Nucleolin, transient receptor potential cation channel, subfamily M, member 2 (TRPM2), and fragment crystallization receptors (FcRs) (see Figure 1).
Astrocytes are the most abundant cell type in the CNS [36]. They build a supportive framework for neuronal networks that helps to mediate blood flow in cerebral tissues and to regulate the allocation of metabolites and neurotransmitters [36,37]. Astrocytes maintain the integrity of the blood–brain barrier and mediate bidirectional transport processes between the CNS and the periphery [38,39]. Their gate-keeping function also comprises the transport of waste products out of the CNS via the glymphatic system and through the blood–brain barrier (BBB) [40,41]. In addition, astrocytes form a complex communication network that exchanges information through the release of signaling molecules [42]. They are in constant cross-talk with neurons, microglia, and oligodendrocytes to uphold a healthy environment in the CNS, and actively participate in neuronal signal transduction and the control of synaptic plasticity [42]. Similar to microglia, astrocytes are able to undergo gliosis to adopt different reactive states with distinct morphology and gene expression profiles [43]. In a pro-inflammatory environment, astrocytes assume their reactive phenotypes where they suspend many supportive functions. Their reactive forms enable them to compensate microglial dysfunction and to engage in phagocytic roles [44]. Accordingly, astrocytes express many pro-inflammatory modulators that are also found on microglia, including Aβ-interaction partners such as TLRs [45], FPRs, and RAGE [46]. Their interaction with Aβ induces severe functional disturbances such as decreased release of neurotrophic and protective factors and a dysregulation of calcium levels, which in turn leads to massive disruptions in gliotransmission and the induction of apoptotic pathways [47]. In addition, astrocytes are able to internalize Aβ and express a number of Aβ-degrading enzymes such as metalloendoproteases like neprilysin (NEP) and insulin-degrading enzyme (IDE) [48], and the matrix metalloproteases MMP-2 and MMP-9 [49], which helps them to support microglia in the clearance of Aβ (reviewed in [50]). Furthermore, several studies have proposed that astrocytes influence Aβ clearance by mediating the transport of Aβ into the CSF and through the BBB into the bloodstream [51,52,53].
Oligodendrocytes are myelinating glial cells of the central nervous system that produce the insulating sheath of axons, which is essential for enhancing axonal action potential conduction [54,55]. In addition, they also contribute to axoglial metabolic support and elimination of oxidative radicals [55]. Not much is known about the involvement of oligodendrocytes in the pathology of AD. However, they express some of the cell-surface receptors that can interact with Aβ such as TLR4 [56,57] and RAGE [58]. Interestingly, recent studies observed molecular heterogeneity of oligodendrocytes in AD mouse models and patients [59,60,61,62,63,64] and three distinct activation states of oligodendrocytes have been identified via single-cell RNA sequencing [65]. Moreover, white matter degeneration and myelin loss have been documented in brains of AD patients [66,67] and studies first theorized that AD might be in part a response to age-related myelin breakdown [68]. Importantly, defects of myelin integrity or demyelinating injuries were recently shown to be drivers of amyloid deposition in vivo [69]. Further detailed investigations are necessary to examine the impact of oligodendrocytes in AD and the contribution of white matter degeneration to AD.

3. Processes That Lead to the Generation of Non-Canonical Aβ Variants

1-42 and Aβ1-40 are by far the best-studied variants in AD research but proteolytic processing of APP generates many more fragments (see Table 1). Thus, in a given biological setting, the different glial cell types and neurons are usually confronted with a much wider spectrum of Aβ variants whose properties and cellular effects are yet insufficiently examined. In general, one can distinguish between four different groups of non-canonical variants: N- and C-terminal abridged peptides, elongated peptides, and peptides containing post-translational modifications (PTMs) (see Figure 2A). A given peptide precursor may be subjected to different cleavage forms and modifications, which gives rise to a wide range of possible combinations of truncations and chemical modification (proteolytic conversion of APP and its modification processes are extensively reviewed in [11]). In addition, different Aβ species tend to form heteromeric aggregates with varying structural properties that may trigger further chemical modifications [12,13]. Several lines of evidence argue that these non-canonical Aβ peptides are of importance in the pathogenesis of AD. Firstly, N-abridged species are the most common form of Aβ peptides and comprise about 70% of all Aβ variants found in human brains [8,70]. Second, their levels in the CSF are directly associated with the onset of AD [71]. Third, some chemical modifications such as pyroglutamylated Aβ, which are dominant in AD but not during normal aging, only occur on abridged peptides [72]. Fourth, some familial APP mutations such as the well-known Swedish, French, or German variants alter APP processing and the ratio of abridged Aβ variants, which may directly influence the progression of AD (see Figure 2B) [73].
Table 1. Overview of abridged and/or modified Aβ-species detected in the brain or CSF of AD patients. Superscript description indicates modification at the respective amino acid (glyco-Y10 = glycosylation at tyrosine at position 10 of the Aβ-domain sequence, ox-M35 = oxidation at methionine at position 35, pyro-E3/11 = pyroglutamylation at glutamate at position 5 or 11, race-D-S26 = racemization of D-conformation serine at position 26).
Table 1. Overview of abridged and/or modified Aβ-species detected in the brain or CSF of AD patients. Superscript description indicates modification at the respective amino acid (glyco-Y10 = glycosylation at tyrosine at position 10 of the Aβ-domain sequence, ox-M35 = oxidation at methionine at position 35, pyro-E3/11 = pyroglutamylation at glutamate at position 5 or 11, race-D-S26 = racemization of D-conformation serine at position 26).
Aβ FragmentsModificationSourceReferences
C-ABRIDGED
1–13 to 1–20-CSF[74,75]
1–15 to 1–20 glyco-Y10glyco-Y10CSF[76]
1–16 to 1–17-Brain, CSF[74,75,77]
1–20-Brain[70]
1–28-CSF[74]
1–30-CSF[74,75]
1–31-Brain[70]
1–33 to 1–34-CSF[74,75]
1–37 to 1–40-Brain, CSF[74,75,78]
1–37 ox-M35 to 1–40ox-M35ox-M35CSF[74]
N-ABRIDGED
2–42 to 11–42-Brain[78,79]
2–40-Brain[79,80]
3–40/42pyro-E3pyro-E3Brain[78,80]
4–42 ox-M35 to 5–42 ox-M35ox-M35Brain[70]
4–40-Brain, CSF[74,78]
4–43-Brain[78]
5–40-Brain[79,80]
8–42 ox-M35ox-M35Brain[70]
9–40-Brain[78]
11–42-Brain[80]
11–42 ox-M35ox-M35Brain[70]
11–42 pyro-E11pyro-E11Brain[78,79]
11–42 pyro-E11, ox-M35pyro-E11, ox-M35Brain[70]
17–42-Brain[81]
C- & N-TRUNCATED
2–14-CSF[75]
2–16-Brain[77]
3–15 to 3–17-Brain[74,77]
3–15 to 4–15 glyco-Y10glyco-Y10CSF[76]
3–19 pyro-E3 to 3–20 pyro-E3pyro-E3Brain[80]
3–24 pyro-E3pyro-E3Brain[80]
4–16 to 5–16-Brain[77]
4–17 to 5–17 glyco-Y10glyco-Y10CSF[76]
4–18 to 4–20-Brain[80]
4–23 to 4 -25-Brain[80]
4–34 -Brain[80]
4–37-Brain[80]
4–37 ox-M35 to 4–40 ox-M35ox-M35Brain[80]
5–20-Brain[80]
11–23 pyro-E11 to 11–25 pyro-E11pyro-E11Brain[80]
11–27 pyro-E11pyro-E11Brain[80]
11–30-CSF[74]
11–34-Brain[70]
25–35/40 race-D-S26race-D-S26Brain[82]
CANONICAL FORMS
1–38 to 1–40-Brain, CSF[74,78,79]
1–42 -Brain, CSF[74,78,79]
1–43-Brain[83]
1–40/42 ox-M35ox-M35Brain[70]
1–40/42race-D-S26race-D-S26Brain, CSF[76,82]
1–40/42 race-D-D7race-D-D7Brain[84,85]
Figure 2. Truncated and modified Aβ species. (A): Depiction of N- and C-truncated Aβ variants and their potential modification sites as described in Kummer and Heneka, 2014. Motifs for β-sheet formation are indicated in red (race = racemization, iso = isomerization, pyro = pyroglutamylation, phos = phosphorylation, glycol = glycosylation, nitra = nitration, ox = oxidation). (B): Aβ peptide with polymorphisms of APP that directly influence the amino acid composition of its Aβ domain (up) [86] and the proposed cleavage sites of various secretases and enzymes (down) [11]. Polymorphisms are classified according to their proposed pathological influence [86]. NEP = neprilysin, PLG = plasmin, ACE = angiotensin converting Enzyme, IDE = insulin-degrading enzyme, M2 = MMP-2, M9 = MMP-9, ECE1/2 = endothelin converting enzyme 1/2, β2 = BACE2, CATB = cathepsin B.
Figure 2. Truncated and modified Aβ species. (A): Depiction of N- and C-truncated Aβ variants and their potential modification sites as described in Kummer and Heneka, 2014. Motifs for β-sheet formation are indicated in red (race = racemization, iso = isomerization, pyro = pyroglutamylation, phos = phosphorylation, glycol = glycosylation, nitra = nitration, ox = oxidation). (B): Aβ peptide with polymorphisms of APP that directly influence the amino acid composition of its Aβ domain (up) [86] and the proposed cleavage sites of various secretases and enzymes (down) [11]. Polymorphisms are classified according to their proposed pathological influence [86]. NEP = neprilysin, PLG = plasmin, ACE = angiotensin converting Enzyme, IDE = insulin-degrading enzyme, M2 = MMP-2, M9 = MMP-9, ECE1/2 = endothelin converting enzyme 1/2, β2 = BACE2, CATB = cathepsin B.
Cells 11 03421 g002
Of note, with respect to these non-canonical Aβ variants, data from the currently available animal models are of limited use since the number of non-canonical Aβ variants in these models diverges significantly from those of most human AD patients, and especially the levels of N-abridged peptides are highly reduced [70,87]. This is likely due to the fact that the Aβ sequence from the most common humanized AD mouse models contain specific mutations of the APP gene or genes of its processing enzymes that are found in some rare familiar cases of AD but not in the majority of patients (Figure 2B) [88,89]. For example, the commonly used APP/PS1 mouse line expresses human APP bearing the Swedish mutation K670N/M671L and mutant human Presenilin-1 [90]. Similarly, the popular 5xFAD mouse model expresses human APP with a non-natural combination of the mutations K670N/M671L (Swedish), I716V (Florida), and V717I (London), and human Presenilin-1 bearing the mutations M146L and L286V [91]. Thus, in comparison to the average human population, these mouse models likely have a bias in APP processing [92]. Nonetheless, humans and transgenic mice share two competing principal pathways for APP processing (see Figure 3). Depending on which cleavage process takes place first, this will lead to different N- and C-terminal amyloid variants that can comprise a length of up to 49 residues [93]. An initial cleavage by α-secretase activity results in short N-terminally abridged peptides starting at position 17, whereas an initial cleavage by β-secretases produces a wide range of Aβ peptides starting at position 1 or 11 [10]. Formally, only peptides that are cleaved at the β-site should be considered as true amyloid β peptides. However, most authors use the “Aβ” abbreviation to also refer to other peptides that are generated through alternative cleavage sites in the amyloid β domain. For simplicity, we will therefore also use this nomenclature, although we are aware that the term amyloid alpha would be more appropriate for fragments that are cleaved by α-secretases. These enzymes produce the membrane-bound α-C-terminal fragment (αCTF) from APP, which is then processed by γ-secretase into amyloid variants comprising the Aβ-domain 17-X, which are also called p3 peptides [94].
β-secretase activity leads to the generation of the classical variants Aβ1-42 and Aβ1-40 mainly via BACE1 (β-site amyloid precursor protein cleaving enzyme 1). BACE1 produces membrane-bound β-C-terminal fragments (β-CTF) by cleaving APP at position 1 (β-site) or at position 11 (β’-site) of the Aβ-domain that are subsequently further processed either by α- or γ-secretases. Cleavage by an α-secretase produces p3 peptides and corresponding N-terminal fragments Aβ1–14 to Aβ1–16. γ-secretase activity is mediated by multiple proteins such as APH-1, PEN-2, Nicastrin, and Presenilin-1 or Presenilin-2 [10]. Peptides that result from γ-secretase activity end with residues ranging between position x-37 to x-43. This process is thought to occur through multiple separate intermediate cleavage steps, where mostly three amino acids are cut off within each step [10]. The cleavage process usually starts at position 48 or 49, continues at position 45 or 46, and ends at position 38, 40, or 42. More rarely, peptides ending at position 34, 37, 39, or 43 are generated [10]. The C-terminal variability is due to the imprecise cleavage of γ-secretases whose recognition motifs are not depending on an exact amino acid sequence. In the healthy brain, 50% of the peptides generated this way end at position 40, 16% at position 38, and 10% at position 42 [99]. These ratios shift in AD, leading to a higher production of peptides ending at position 42 [10]. The reason for this is not fully understood; however, mutations of either APP or γ-secretase enzyme complexes can lead to a lowered binding of Aβ to the protease complex, which may then result in a premature release of the cleaved peptides [10]. In addition to these longer variants, γ-secretase can also generate shorter peptides (Aβ1–14 to Aβ1–16); however, the exact pathways have not been identified, yet [100]. In addition to processing by traditional secretases, Aβ and APP are targeted by a wide range of enzymes and peptidases that can either cut APP directly or interact with Aβ peptides after their generation (see Figure 2B) (reviewed in [11]). These include matrix metalloproteinases such as MMP2 and MMP9, which mainly interact with extracellular or internalized Aβ in glial cells or deposition of fibrillary Aβ aggregates or others such as meprin-β that cleave APP directly in its membrane bound form [11]. This leads to an extensive repertoire of possible peptides (see Table 1).

3.1. N-Abridged Aβ Species

2-X and 3-X are thought to be generated by APP cleavage by the metalloprotease meprin-β or after cleavage of “full-length” Aβ1-X by the exopeptidase aminopeptidase A [101]. Not much is known about the differences between Aβ2-X and non-abridged variants regarding their biological effects. However, Aβ2-40 shows a higher aggregation propensity than Aβ1-40 and can seed further aggregate formation of Aβ1-40 [102]. Cleavage by meprin-β depends on a specific recognition motif in APP that includes amino acids neighboring the N-terminus of Aβ [102]. This motif is altered in some APP mutants such as the Swedish APP variant, which leads to a diminished processing by these enzymes and a lower number of these variants in patients that carry this mutation [102]. Of note, many popular humanized AD mouse models such as 5xFAD and APP/PS1 mice express the Swedish APP mutation [92]. Thus, these mice have a low amount of Aβ2-X. By contrast, these N-terminal truncations are frequent in human AD patients [103].
4-X variants are abundantly found in dense amyloid plaque cores of human AD patients and 5xFAD mice [104]. Interestingly, Aβ4-X peptides appear to be more commonly released by microglia and astrocytes than by neurons, but the exact processing mechanism is currently unknown [98]. Several studies suggest that a disintegrin and metalloproteinase with thrombospondin motifs (ADAMTS4) and neprilysin may generate these truncations [104,105,106]. ADAMTS4 has been implicated in the generation of variants such as Aβ4-40 and Aβ4-42 [105]. Both show increased aggregation rates and a similar neurotoxicity as non-abridged Aβ [95,107]. Clearance of oligomeric Aβ4-40 and Aβ4-42 from the brain is markedly reduced compared to their non-abridged counterparts [108], which may indicate reduced or hampered interactions with glial cells. Of note, injections of Aβ4-40 or Aβ4-42 into the brain of mice elicit behavioral deficits and impaired memory, which were comparable with those induced by Aβ1-42, while injections with Aβ4-38 had no effect [95]. Interestingly, transgenic mice expressing human Aβ4-42 in their hippocampus do not develop plaques but still suffer from hippocampal pyramidal neuron loss and behavioral deficits [95,109].
11-X peptides are generated by the β-secretase processing pathway after cleavage at the β’-site. Aβ11-40 and Aβ11-42 levels are commonly elevated in the brain of AD patients [70,79]. They are approximately as abundant as Aβ1-42 and can comprise up to 20% of the peptide content in plaques [70,79]. Decreased levels of Aβ11-X in the CSF are associated with the onset of symptoms in AD and have thus been proposed as a biomarker for AD [71]. In plaque cores, Aβ11-X co-localizes with Aβ1-42 and is therefore believed to contribute to early plaque formation [110]. In vitro studies demonstrated that Aβ11-40 leads to increased neurotoxicity exceeding that of the canonical Aβ1-40 [111]. Furthermore, Aβ11-40 is capable of rapidly forming fibrils [112] and can therefore act as a cross-seed for the deposition of other Aβ species in plaques [111]. Interestingly, homogenous aggregation of Aβ11-40 decreases its neurotoxicity while heterogeneous aggregates of Aβ11-40 and Aβ1-40 are even more harmful than monomeric or homogeneous oligomers of Aβ1-40 [111]. Surprisingly little is known about the effects of Aβ11-40 on glial cells. However, we recently showed that Aβ11-40 activates the human glial cell line U87 via FPR1 [113].
17-40 and 17-42 are p3 peptides that are generated via the α-secretase processing pathway. P3 peptides are found in diffuse plaques of AD patients [114] and microglia of AD patients contain various N-terminally abridged Aβ species [115]. Cell culture experiments indicate that they might be produced in even higher amounts than Aβ1-42 and Aβ1-40 [116]. Of note, the presence of Aβ17-X in the CSF is positively correlated with cognitive impairment of AD patients [71], which argues for a contribution of these peptides to AD. However, they are difficult to detect because many commonly used Aβ antibodies are not selective for Aβ1-42 and cross-react with p3 fragments and other N-terminally abridged variants [117]. In addition, many established protocols for isolating Aβ peptides from brain and CSF samples are not suitable for p3 peptides due to their high hydrophobicity and insolubility, which excluded them from many studies that evaluated Aβ levels [118]. Moreover, since several non-canonical proteases such as IDE, NEP and MMP9 have cleavage motifs within the p3 sequence (Figure 2B), it is likely that additional p3 peptides with highly variable C-termini exist. However, these potential variants have not been studied so far. Historically, Aβ17-X were classified as non-amyloidogenic because they did not display fibrillary aggregation and deposition. However, recent advances have cast doubt on these assumptions because conflicting data regarding the amyloidogenic capabilities of p3 peptides exist (reviewed in [94]). While early reports found no evidence of toxic fibril formation by p3 peptides [119], more recent studies found that p3 peptides form fibrils and soluble oligomers [120,121,122] with even faster aggregation kinetics than Aβ1-40 or Aβ1-42 [107,120]. Aβ17-40 and Aβ17-42 induce pro-inflammatory cytokine production in human and murine glial cell lines, leading to neuronal decay through pro-inflammatory caspase-activation in neuronal cell lines [107,123], which triggers pro-inflammatory activity after injection into mouse brains [124]. Moreover, we recently showed that all human FPRs can detect Aβ17-40 and that FPR1 is activated at 30-fold lower concentrations by the peptide than by Aβ1-42 [113].

3.2. C-Terminal Variants

1–43 is produced by γ-secretase cleavage and can be found abundantly in amyloid plaques [125,126,127]. Its levels are heightened in both spontaneous and familiar AD patients, but not in age-matched controls [128]. In particular, familiar AD cases with mutations in the γ-secretase subunit PS1 show an elevated generation of Aβ1–43 compared to other species [129,130]. Unlike shorter variants, only low amounts of Aβ1–43 are found in the CSF [131] and cerebral blood vessels [125]. Aβ1–43 is highly amyloidogenic and neurotoxic in mouse neuronal primary cultures and cell lines [132,133]. Furthermore, injections of Aβ1–43 into APP mice lead to severe depositions of Aβ1–42, indicating the potent capability of Aβ1–43 to seed the aggregation of other variants [134].
1–37 to 1–39 are produced by the traditional β-secretase pathway. Compared to canonical Aβ1–42 and Aβ1–40, these C-abridged peptides appear to be harmless or at least significantly less neurotoxic [135,136]. Aβ1–38 forms oligomers and fibrils, but its low abundance in plaques indicates that it is easily cleared from the CNS. Furthermore, levels of these C-abridged peptides in CSF are altered in patients with AD [137]. Functionally, C-abridged Aβ species can act as a scavenger for Aβ1–42 and Aβ1–40 because they can form aggregates with these peptides, leading to reduced synaptic disruption, reduced neurotoxicity, and lowered amyloidogenicity [136]. In a Drosophila melanogaster model, expression of Aβ1–42 led to severe neuronal and behavioral pathologies, but co-expression of either Aβ1–37,1–38 or Aβ1–39 attenuated these effects [135]. In addition, a lower ratio of soluble Aβ1–42/Aβ1–38 (i.e., a higher Aβ1–38 content) was associated with a later age-at-death in male AD patients [136].
1–34 is present in the brain and CSF of AD patients and AD mouse models [138,139]. Its levels are highly increased in the CSF of AD patients compared to healthy controls [139]. Interestingly, Aβ1–34 is also commonly found in brain vessels in early AD stages but diminishes with further progression of the disease [139]. Aβ1–34 generation processes have not been fully elucidated. In cell culture experiments and transgenic mice, generation of Aβ1–34 depended on initial BACE1 activity [139], but its C-terminal truncation is likely produced through procession of Aβ1–40 or Aβ1–40 by glia-derived metalloproteases such as MMP2 and MMP9 [140]. Thus, Aβ1–34 is thought to be a degradation marker for clearance activity of microglia in early stages of AD [138,139,140]. Interestingly, an in vitro study reported that, Aβ1–34 protected APP-expressing HEK293 cells against caspase-3-mediated apoptosis, which may indicate that Aβ1–34 is not just a side product of degradation processes but may also have resolving properties [141].
1–14 to 1–16 are commonly generated during APP processing but are thought to be harmless [142] because they do not elicit pro-inflammatory glia signaling, are not capable to form amyloidogenic aggregates, and are most likely quickly degraded. Increased levels of these fragments in the CSF indicate increased β-secretase activity and have thus been proposed as a biomarker for Aβ production [78,100].

3.3. Post-Translational Aβ Modifications (PTMs)

Pyroglutamylation refers to the conversion of glutamate into pyroglutamate, which can occur in N-abridged Aβ peptides at position 3 (Aβ3-xpyro-E3) and position 11 (Aβ11-xpyroE11). This process was shown to be catalyzed by glutaminyl cyclase in vitro [143,144] and also in vivo [145,146]. Pyroglutamate highly influences the secondary structure of Aβ, leading to an enhanced β-sheet structure, higher hydrophobicity, and increased aggregation propensity [147]. AβpyroE species are peculiarly abundant in senile plaques [78,148] and are also present in diffuse, soluble aggregates [149,150]. Aβ3-Xpyro-E3 is highly amyloidogenic and can seed the aggregation of other Aβ species, which increases fibril formation significantly [151,152]. Such cross-seeded aggregates appear to be more neurotoxic than homogenic aggregates and may thus contribute to the pathological effects of AD [152]. Unlike most Aβ variants, AβpyroE is only associated with the progression of AD but not with ageing [72]. In pre-clinical AD patients without senile plaque formation, Aβpyro is not initially present within soluble aggregates [149] but appears during further progression of the disease when patients experience their first cognitive symptoms [149,153]. Deposition of Aβ3-40 pyro-E3 into insoluble plaques precedes deposition of Aβ1–42 and Aβ1–40 [148] and may thus be involved during the onset of early symptoms. Aβ3-X pyro-E3 species also accumulate in lysosomes of microglia [154] where they are suspected to cause lysosomal failure, microglial decay, and subsequent progression of AD [155,156]. In addition, pyroglutamylation may affect protease-mediated decay, and influence the capabilities of Aβ to interact with cell-surface receptors. For example, murine NMDAR recognizes Aβ but not Aβ3-X pyro-E3 [157]. Furthermore, experiments with porcine primary microglia revealed that Aβ3–42pyro-E3 is more potent to enhance E. coli phagocytosis than the non-abridged peptide (360). Unfortunately, production and deposition of AβpyroE in most AD animal models is not comparable with human AD patients because in APP transgenic mouse lines initial plaques are devoid of Aβ3-X pyro-E3 species [158,159,160]. Nonetheless, in some cases mouse models can provide valuable insights into the biological effects of AβpyroE species. For example, 5xFAD mice overexpressing the human glutaminyl cyclase show an increased AβpyroE production, plaque formation, and behavioral symptoms, while the glutaminyl cyclase knockout reduces plaque deposition and rescues behavioral performance [145]. Moreover, treatment of APP/PS1 mice with antibodies against Aβ3-Xpyro-E3 ameliorates behavioral symptoms and decreases amyloid plaque numbers, which is likely due to FcR-mediated clearance by microglia [161].
Isomerization of Aβ occurs mainly at asparagine (N) and aspartate (D) residues and is commonly detected at position 1 (Aβiso-D1) and 7 (Aβiso-D7) [162]. Isomerization happens through spontaneous, non-enzymatic reactions. Thus, its probability increases throughout the lifetime of Aβ [163]. In plaque core preparations, Aβ species with iso-D1 and iso-D7 are more common than other variants and are present in higher concentrations than in “younger” diffuse plaques or vascular depositions [164,165]. Of note, isomerization has a strong impact on the secondary structure of Aβ and leads to faster aggregation and deposition [166,167]. Aβiso-D7 also shows an increased neurotoxicity through excessive NO generation [168,169] and inhibits the α7 Nicotinic receptor, which has been implicated in long-term memory formation [168]. Microglia can internalize isomeric Aβ [164]; however, isomeric Aβ is more resistant against degradation in microglial lysozymes [163] and against intracellular and extracellular proteases [163,170], which fosters lysosomal failure [155,156].
Phosphorylation can potentially occur at serine residues 8 (Aβpho-S8) and 26 (Aβpho-S26) and at tyrosine residue 10 (Aβpho-T10) [13]. These modifications are thought to occur mainly through extracellular kinases or after internalization. pho-S26 accumulates in early AD stages in neurons, but only low concentrations are present in extracellular plaques [171,172]. It can assemble into soluble oligomers that do not form fibrils and exhibit increased neurotoxicity in cell culture experiments [172]. pho-S8 levels are elevated in later stages of AD where it mainly occurs in compact plaques [173]. A phosphorylated serine at position 8 increases the stability of Aβpho-S8 aggregates by attenuating their recognition by degrading enzymes and inhibiting microglial clearance [174,175]. Due to the clearance resistance, Aβpho-S8 species may contribute to plaque spreading [175]. In this context, Hu and colleagues showed that Aβ1-42pho-S8 can cross-seed with non-modified Aβ1-42, which generates aggregates with elevated neurotoxicity [152].
Oxidation is mainly caused by radicals such as ROS and NO and can occur at the methionine (M) at position 35 (Aβox-M35). Aβox-M35 molecules are abundantly released by reactive microglia. Oxidation disrupts fibril formation and destabilizes oligomer formation [176,177]. Since generation of oxidative stress is a hallmark of neuroinflammation, the amount of oxidized Aβ is thought to increase during the progression of AD. Oxidized Aβ species with varying length have been identified in brain and CSF samples of AD patients [70,74,80]. A study by Head and colleagues found oxidized species in 46% of diffuse plaques and in 98% of cored plaques [178]. Oxidized Aβ was found in plaque-invading microglia of AD patients [178]. However, so far it has not been elucidated if oxidation influences interactions between Aβ and microglia.
Racemization converts amino acids from their L- into D-conformation. For Aβ, racemization was observed in aspartate residues (D) at position 1 (AβD-D1) and in seryl residues (S) at position 26 (AβD-S26) [13]. Racemization of aspartic acid increases the fibrillary aggregation kinetic [179]. In addition, further modifications such as sylation and nitration were observed [13]. However, their biological relevance is yet unclear. Some studies suggest that these PTMs may also influence the aggregation properties of Aβ and could influence cross-seeding with other Aβ species [152].

3.4. Splice Variants of APP

The APP gene is highly conserved in mammals, consists of 18 exons, and spans over 290 kilobases [89,180]. Ten APP splice variants have been described that range between 639 to 770 amino acids [89]. The three variants APP695, APP751, and APP770 make up almost the complete protein quantity [181,182]. APP770 is the longest isoform and contains all exons [181,182]. In comparison, APP751 does not include exon 7, which encodes for an additional Kunitz-type protease inhibitor (KPI) domain that protects the protein against proteolytic cleavage by certain enzymes [183]. APP695 lacks both exons 7 and 8,and is therefore deficient of the KPI domain from exon 7 and an OX-2 domain from exon 8, which is thought to influence cell-surface binding [184]. All three mayor isoforms contain the full Aβ sequence and can thus potentially be processed into Aβ and p3 peptides. Interestingly, APP695 is the most common isoform in the brain and is mostly expressed by neurons, whereas APP751 and APP770 are more dominant in glial cells [181,182,184].

3.5. APP Processing Is Different in Neurons, Astrocytes, Microglia, and Oligodendrocytes

Neurons are beyond doubt the primary source for the canonical Aβ1-40 and Aβ1-42 variants [185,186]. They mostly express the APP isoform APP695 and only minor amounts of APP751 and APP770 [187,188]. Neuronal mRNA levels of APP are approximately 10-fold higher [189] than in other cell types and neurons produce approximately four-fold more soluble APP proteins than astrocytes and microglia [190]. In line with this, Aβ production was also shown to be elevated in neuronal cultures compared to astrocytes or microglia cultures [96]. This is corroborated by a study reporting a ~7× higher generation of Aβ1-40 secreted by neuronal cultures compared to Aβ released from astrocytic cultures [98]. Neurons predominantly secrete Aβ peptides starting at position 1 (80%), whereas those released from astrocytes and microglia represent mainly N-terminally abridged Aβ peptides including Aβ2/3-x and Aβ4/5-x (Figure 3B) [98], suggesting that secretase levels are different in the respective cell types. In line with this, BACE1 protein levels were reported to be more abundant in cultured neurons compared to astrocytes, whereas BACE2 showed a higher presence in astrocytes [191]. Furthermore, neurons produce 3.4 times more Aβ17-x than Aβ1-40-43, while astrocytes generate 7.8 times more Aβ17-x than Aβ1-40-43, indicating a higher rate of α-secretase cleavage in astrocytes compared to neurons [96]. Protein expression of PS1, the catalytic domain of the γ-secretase complex, was shown to be comparable in neurons and astrocytes, suggesting no major difference regarding γ-secretase processing in both cell types [191].
In astrocytes, all three major splice forms of APP, APP695, APP751, and APP770, were shown on the protein level [192]. Splice forms APP751 and APP770 seem to be predominant [190,192], while APP695, which is the major described APP splice form in neurons [187,188], has a lower abundance. This might contribute to differences regarding the proteolytic conversion of APP between astrocytes and neurons. In contrast to neurons, only 40% of the secreted Aβ peptides are cleaved at position 1 in astrocytes [98]. This is of interest, because N-terminally abridged Aβ species are prevalent in neuritic plaques [77,95,193,194]. N-terminally abridged Aβ species from astrocytes and microglia may therefore contribute to a higher proportion to the formation of neuritic plaques than peptides derived from neurons. In line with this idea, cell culture experiments demonstrated that astrocytes and microglia indeed secrete higher amounts of N-terminally Aβ variants such as Aβ2/3-40 and Aβ4/5-40 [98]. Proteolytic conversion of these variants does not depend on BACE1 activity but is likely mediated by plasma-membrane associated cathepsin B (CatB) [195]. However, a number of other proteases such as meprin β [196], neprilysin [197], myelin basic protein [198], the metalloproteinase ADAMTS4 [105], and aminopeptidases [199] may also contribute to the production of these N-terminally modified Aβ variants [11,200]. By contrast, the proportions of the C-terminal Aβ variants Aβ1-37, Aβ1-38, Aβ1-39, and Aβ1-42 to Aβ1-40 did not differ between neurons, astrocytes, and microglia [98], indicating no differences in the C-terminal Aβ peptide-trimming by γ-secretase between these different cell types [10]. Cortical astrocyte cultures have lower BACE1 protein levels and higher levels of BACE2 than corresponding neuron cultures [191]. In line with the finding that APP is processed by BACE2 at position 19–20 and 21–22, Western blot experiments demonstrated a more efficient generation of Aβ1-15, Aβ1-19, and Aβ1-20 in astrocytes of APP/PS1 mice [191]. However, the isolation of primary astrocytes typically yields cultures with mostly reactive phenotypes [201], and these therefore possess presumably higher APP protein levels than resting astrocytes since it has been shown that APP expression in reactive astrocytes is substantially increased following neuronal damage [202]. By contrast, only little APP expression is documented in resting astrocytes in vivo [203]. Therefore, primary cultures may not exactly reflect the in vivo situation of the cell types analyzed. Of note, BACE1 protein and APP levels can be induced by 300 to 600% through stimulation of astrocytes by cytokine combinations or Aβ1-42 [204].
In microglia, APP protein expression was reported at an early stage [205,206]. All three major splice forms, APP695, APP751, and APP770, are present in microglia cell cultures [192]. However, microglia express more than 50% of their APP mRNA as transcripts as APP695, which encoded for the KPI-domain; approximately 22% of total APP mRNA represented APP770 mRNA, 45% APP751/L-APP752 mRNA, 25% L-APP733 mRNA, and approximately 4% APP695 and L-APP677 [207]. Furthermore, BACE1 expression has been indicated on a protein as well as the mRNA level in microglia cells [208]. The presence of α-secretase ADAM10 in microglia was also shown via an upregulation of ADAM10 after reduction in cortical activity [209]. Interestingly, protein amounts of gamma secretase subunits such as PS1 and Nicastrin are increased in activated microglia [210].
The whole APP processing machinery is present in microglia under certain conditions. However, the impact of microglia on APP processing has not been characterized extensively, yet. Microglia produce mainly N-terminally modified Aβ peptides including Aβ2/3-x and Aβ4/5-x [98], suggesting a possible contribution of N-abridged Aβ species from microglia to the peptides found in vascular and neuritic plaques [77,194]. Several studies with the immortalized mouse microglial cell line BV-2 revealed an interaction of microglia with the extracellular matrix that affects APP secretion as well as the intracellular biogenesis of APP [207]. APP expression and release of soluble APP was highest after adherence to uncoated plastic surfaces [207]. A further study using this cell line showed that Aβ25-35 and lipopolysaccharide treatment induced Aβ secretion [211]. APP is expressed on the surface of microglia cells and it is upregulated due to their activation [205,206]. Direct stimulation of APP with agonist antibodies also led to robust activation of microglia [212,213]. Even an activation of microglia via Aβ-binding to APP has been indicated [214]. Collectively, these data indicate that APP itself can serve as a cell-surface receptor that helps regulating microglial activation.
In oligodendrocytes, analysis of mRNA transcripts identified the APP splice forms APP695, APP751, and APP770 [215]. Interestingly, a lower molecular weight form of full-length APP was more substantially expressed than higher molecular weight forms, suggesting that oligodendrocytes, similar to neurons [216], may express more APP695 than APP751 and APP770 [217]. Adult oligodendrocyte progenitor cells (aOPCs), which were purified from a >4-month-old rat brain, generally expressed BACE1, APP, and components of the γ-secretase complex such as Presenilin-1, anterior pharynx-defective 1 (APH1), and Nicastrin [218]. APP protein expression in oligodendrocytes has also been shown via immunostaining of rodent brain [219,220,221], as well as in cell culture [215,217,222]. Moreover, ADAM10 expression has been documented in oligodendrocytes of the developing brain at later embryonic stages [223]. APP processing analyses in a cell line of murine OPCs (Oli-neu) [224] documented generation of the APP intracellular domain (AICD) by γ-secretase and production of an APP C-terminal fragment, which increases after treatment with a γ-secretase inhibitor [225]. A more detailed study even addressed various differentiation stages of cultured aOPCs and APP processing [218]. These aOPCs can be maintained for several months in media containing Fibroblast growth factor 2 (FGF2) which promotes survival and proliferation of aOPCs positive for NG2 (chondroitin sulfate proteoglycyan NG2/CSPG4), a marker for aOPCs [218]. Interestingly, FGF2 withdrawal increased expression of full-length APP as well as Aβ1-42 production. aOPC cultures including only 0 or 1 ng/mL FGF2 secreted ~four-fold higher ratios of Aβx-42 to total Aβ than cultured fetal rat neurons, suggesting a high rate of γ-secretase processing in these aOPC cultures [218]. α-CTF accumulated in media without FGF2 while β-CTF and the p3 fragment increased with elevated FGF2 concentrations, suggesting that γ-secretase prefers the APP α-CTF as a substrate at higher concentrations of FGF2. Further proteases that are capable of processing Aβ have also been found in oligodendrocytes. One example is ADAMTS4, which is exclusively expressed in oligodendrocytes in mouse brain and generates Aβ4-x from both full-length APP and Aβ1-x [105]. Mass spectrometric profiles of different Aβ species from 5-day-old OPC cultures from WT and ADAMTS4 KO mice confirmed release of Aβ4-x species from these cells [105]. Interestingly, ADAMTS4 mRNA levels were increasing during the culture period of 5 days, suggesting a higher conversion of APP by ADAMTS4 in myelinating oligodendrocytes [105]. Oligodendrocytes therefore seem to be a source of secreted Aβ4-x peptides [105], which corresponds to the finding of N-terminally abridged Aβ variants such as Aβ2/3-40 and Aβ4/5-40 secreted by microglia and astrocytes in contrast to neurons [98].

4. Cellular Responses of Glial Cells towards Aβ

4.1. Contribution of Aβ Internalization and Degradation

Microglia help to remove toxic peptide Aβ isoforms from the CNS via internalization and subsequent degradation by several mechanisms (see Figure 4) [226]. Various microglial receptors are able to mediate uptake of Aβ by endocytosis via formation of receptor-peptide complexes. However, different aggregates seem to preferentially interact with different types of cell-surface receptors. For example, FPRs help in the clearance of Aβ1-42 monomers and small oligomers [227,228,229], whereas fibrillary aggregates are not efficiently internalized [230,231]. Scavenger receptors such as SR-A [232], SR-B1 [233], and CD36 [234,235] appear to prefer the uptake of fibrillary forms of Aβ1-42, whereas TLRs have been reported to mediate the internalization of monomeric, oligomeric, and fibrillary Aβ1-42 [226]. TLR2 [236,237] and TLR4 [238] induce the uptake of Aβ1-42 and trigger internal degradation processes. In addition, microglia can also internalize Aβ1-42 via unspecific uptake mechanisms, such as pinocytosis through phosphoinositid-3-kinase (PI3K) [239] or Ras-related C3 botulinum toxin substrate 1 (Rac1)-dependent pathways [240]. Internalization leads to different internal degradation processes in microglia. In cell culture experiments, fibrillar Aβ1-42 was effectively degraded through autophagy by lysosomes [241,242,243]. The acidic environment in lysosomes enables the formation of compact Aβ1-42 aggregates that are resistant against further degradation [155,156]. Microglia can release these aggregates as microvesicles into the extracellular space where they contribute to plaque formation and neurotoxic effects [155,156]. This can trigger a vicious cycle of apoptotic pathways [155] involving an aberrant activation of the NLRP3 inflammasome [244] and the apoptosis of microglia [155]. Some studies suggest that the release of degradation-resistant Aβ1-42 from dying microglia might be the driving force behind plaque deposition since microglia-deficient animal models show only minimal Aβ1-42 plaque formation [245,246]. Interestingly, autophagy of Aβ in microglia seems to become less efficient with disease progression [247].

4.2. Contribution of NLRP3 Signaling

The wide range of interaction partners for Aβ leads to a number of equally diverse and complex interactions between different signaling cascades in the inflammatory response of microglia (see Figure 4 and Figure 5) [3,23]. These cascades are mainly regulated by inflammasomes, which are cytosolic protein complexes that mediate inflammatory responses of immune cells [249]. In AD, formation of the NLRP3 inflammasome and subsequent caspase-1 activity have been proposed to be driving forces in the microglial response to Aβ [248]. The main components of this inflammasome are the cytosolic receptor NLRP3, the adaptor protein adaptor protein apoptosis-associated speck-like protein containing a caspase activation and recruitment domain (ASC), and the inactive precursor form of caspase-1 (pro-caspase-1) [248,250]. Priming and assembly of the NLRP3 inflammasome is initiated by an activation of NF-κB pathways that are typically induced by extracellular stimulation via surface receptors [248] such as TLRs [236], CD36 [251], and FPRs [252]. This leads to NLRP3 upregulation, which then recruits ASC and pro-caspase-1 for inflammasome formation [3,248]. Upon a second signal, which is either triggered by cell-surface receptors or through direct interaction of NLRP3 with internalized Aβ [253], the primed inflammasome complex autocatalyzes pro-caspase-1 through the adaptor protein ASC into its active state, which subsequently triggers transformation of interleukines and cytokines into their biologically active forms [248]. Of note, microglia release inflammasome components such as ASC specks into the extracellular space, where they can be a seed for Aβ, which may propagate plaque formation [254]. In general, detection of Aβ induces the release of Il1β, IL6, IL8, IL33, IL34, and other inflammatory factors such as TNFα. IL1 and its isoform Il1β belong to the most prominent representatives because upregulation of Il1β already occurs at the earliest stages of plaque evolution [255]. Il1β release increases APP production in neurons [256,257] and influences APP processing via upregulation of proteases [258,259,260]. High IL1 levels were found in brain tissue [261,262] and cerebrospinal fluid [261] of AD patients and APP/PS1 mice [262]. Il1β induces further self-production in microglia and astrocytes, leading to a self-potentiating inflammatory feedback loop [263]. This is associated with neuronal toxicity, exacerbated synaptic loss and aberrant glial activation.

4.3. Contribution of Oxidative Stress

Generation of ROS and radical nitrogen species (RNS) is a typical pro-inflammatory action of microglia to remove extracellular pathogens [3,264]. Microglia can release an oxidative burst, which is usually driven by the transmembrane enzyme complex NADPH oxidase (NOX) [264]. NOX activity is triggered as a secondary effect after the phagocytosis of pathogens or the detection of PAMPS or DAMPS by cell-surface receptors such as TREM2 [265] and FPRs [266]. This leads to the generation and release of ROS into the extracellular space. Unfortunately, this does not only target harmful pathogens but also harms neighboring cells, beneficial proteins, and lipids [264]. The healthy brain combats oxidative stress through a complex antioxidant system, which includes radical scavengers such as glutathione that protect cell organelles from damage and dysfunction [264,267]. However, aberrant release of ROS and RNS can overwhelm this antioxidant defense system, leading to severe damage through lipid peroxidation, protein oxidation, and DNA damage [264].
Generation of oxidative stress has been described as a hallmark feature of AD [7]. Signs of oxidative damage are commonly found in the brain, CSF, blood, and urine of patients [268,269,270,271]. In particular, brain areas afflicted with large plaque and NFT depositions show high levels of oxidated proteins and lipid peroxidation. Next, regulatory factors of NOX are upregulated in AD brain tissue [272,273]. Aβ triggers ROS release through activation of microglial cell-surface receptors. The ROS effects are initially dampened by the antioxidant system, but chronic exposure of cells to Aβ leads to its depletion and thus the subsequent exacerbation of oxidative damage [274]. Aβ-mediated oxidative stress then disrupts synaptic signaling, induces astrocyte activation, and leads with increasing levels to neuronal death. In addition, ROS activates TRPM2, which propagates the pro-inflammatory activity of microglia and initiates apoptotic pathways in neurons [275]. Apoptotic cells subsequently release additional DAMPs that are perceived by microglia and further stimulate the pro-inflammatory machinery, thus leading to the self-renewal of oxidative stress. Moreover, oxidative stress leads to increased production of Aβ through modulation of APP procession pathways [276,277,278,279]. Treatment of neuronal cell lines with H2O2 increases expression and activity of β- and γ-secretases that mediate the production of pathogenic Aβ species [276,278,279]. In particular, γ-secretase activity seems to be susceptible to the effects of oxidative stress and may result in increased β-secretase levels through yet unknown pathways [279,280]. In addition, oxidative stress can directly facilitate the generation of modified Aβ species through oxidation at position 35 or nitrosylation at position 10 [13].

4.4. Interaction of Aβ with Other Neuropathological Proteins

Amylin (also known as islet amyloid polypeptide (IAPP)) is a peptide hormone generated by islet β cells in the pancreas that is co-secreted with insulin into the blood stream during regulation of blood glucose levels [281]. Amylin-Aβ aggregates are present in cerebral plaques of patients with familiar AD [282]. In CSF samples of familiar AD patients, high Amylin concentrations are associated with decreased Aβ1-42 levels, which suggests that Amylin may influence Aβ1-42 transport between the brain, CSF, and blood [282,283]. Interestingly, the Amylin receptor may also be able to interact with Aβ [284]. Amylin and Aβ1-42 share a sequence homology of more than 50% and form similar β-sheet structures. Moreover, Amylin can act as a seed for Aβ1-42, which can lead to amorphous heterocomplexes that display a highly increased neurotoxicity compared to either pure Amylin and Aβ1-42 aggregates [285].
Prion protein (PrP) is a cell-surface protein encoded by the PRNP gene, which is ubiquitously expressed throughout human tissues, but is most abundant in the brain (reviewed in [286]). Here, PrP is mainly expressed in astrocytes and neurons, but low levels are also present in microglia and oligodendrocytes [287,288,289]. In addition to its membrane-anchored isoform, soluble PrP variants have been described [290,291]. PrP’s biological functions are not fully elucidated but several studies argue for PrP contribution to synapse formation and neuronal signal transduction [292,293,294]. Its physiological form is dominated by an α-helical secondary structure and is referred to as PrPc (cellular form) [286]. However, under pathological condition, PrP undergoes a structural shift into a neurotoxic isoform with predominant β-sheet conformation, also called PrPsc (scrapie form) [295]. Formation of PrPsc is thought to be contagious, leading to further conversion of PrPc into their pathogenic form [296]. PrPsc form aggregates with itself and other proteins that are highly neurotoxic and lead to severe disruption of normal CNS functions [295]. PrP is commonly found in the diffuse and dense plaques of AD patients and where it can be co-localized with Aβ [297,298,299]. In addition, PrP can directly bind Aβ in both its soluble and membrane-bound form [300,301]. However, the biological consequences of these interactions are still debated because deletion of the PRNP gene did not ameliorate plaque formation in APP/PS1 mice but ameliorated behavioral and cognitive symptoms [302,303]. Interestingly, PrP was shown to modulate BACE1 activity and thus the production of Aβ, but had only diminished effects on processing of the Swedish APP variant [304,305].

5. Genetic Variants Affecting Microglial Response to Aβ

Microglia and other glial cells express various receptors and molecules that can either directly bind Aβ at the cell surface, which leads to altered cell signaling, or interact with it indirectly, which modifies Aβ detection or degradation (Figure 5).
TREM2 is a cell-surface receptor that is highly expressed in microglia. It mostly detects phospholipids and glycolipids that are released from host-derived apoptotic cells or from invading pathogens (Yeh 2017), but is also able to interact with multiple peptides and proteins such as Aβ and various Apolipoproteins (Apo) including APOE and APOA1 [306]. TREM2 is capable of inducing cell activation and migration towards Aβ1-42 [307]. In TREM2-deficient AD mouse models such as modified APP/PS1 and 5xFAD, microglia do not congregate around senile plaques and thus do not respond to the progressing plaque load [306]. In recent years, this receptor has received much attention, since loss of TREM2 leads to nearly complete absence of microglial reactivity in mice [308,309]. In addition, several genome-wide association studies (GWASs) in humans found mutations and polymorphisms of the TREM2 gene to be a high-risk factor for familiar forms of AD (see Table 2).
TLRs comprise a family of pattern recognition receptors (PRRs) that detect pathogen-associated molecular patterns (PAMPs) from exogenous sources such as bacteria and fungi, but can also recognize damage-associated molecular patterns (DAMPs) from endogenous sources [355]. There are ten receptor subtypes (TLR1–10) in humans that are all expressed in the central nervous system (CNS) as well as in other tissues [355,356]. In microglia, TLRs induce cell activation after detection of harmful stimuli and initiate the inflammatory cascade which includes the upregulation of additional PRRs and the production of cytokines and reactive oxygen species (ROS) [357]. In the context of AD, TLR2 and TLR4 are of special interest since they are capable to detect Aβ1-42. Both TLR2 and TLR4 mediate the phagocytic uptake and degradation of Aβ in microglia [236,358] and induce typical pro-inflammatory cascades such as the release of Il1β and TNFα as a response to the peptides [359]. Their expression is upregulated in inflamed brain tissue and in plaque-invading microglia of human AD patients [359] but also in microglia derived from APP/PS1 mice [238]. APP/PS1 mice with dysfunctional TLR4 display decreased cytokine levels [360] but also more severe plaque formation [238]. Similarly, TLR2-deficient APP/PS1 animals show accelerated cognitive impairment and increased Aβ1-42 concentrations in the brain [237]. On the contrary, in a study with APP/PS1 mice, long-term administration of TLR2-inhibiting antibodies decreased microglial activity, improved cognitive performance, and lowered Aβ plaque load [361].
CD36 is a class B scavenger receptor that is expressed in microglia, astrocytes, and neurons in the CNS, and in various peripheral cell types such as innate immune cells, myocytes, and endothelial cells [362]. CD36 is mostly known as a fatty acid transporter but is also involved in lipid metabolism and the regulation of inflammatory responses in immune cells [363]. Its ligands include various structural proteins of the extracellular matrix such as collagens and thrombospondins. In addition, CD36 interacts with fibrillary Aβ, which induces a pro-inflammatory activation of microglia [235,364]. Upon detection of Aβ, CD36 mediates cell migration and subsequent binding to Aβ fibrils [364]. It contributes to cytokine production [251] and generation of oxidative stress [234]. In addition, CD36 may act as a co-receptor for other PRRs that interact with Aβ. For example, activation of CD36 can lead to the formation of complexes with TLR2 and TLR6, thus eliciting internalization and degradation of Aβ1-42 [365]. Of note, in human brain samples, expression of CD36 is detected in microglia proximal to Aβ plaques [366]. However, Ricciarelli and colleagues reported that CD36 is also highly expressed in healthy people with senile plaques [366].
RAGE is a receptor mainly binding advanced glycation endproducts, which are generated by non-enzymatic glycation and oxidation of proteins and lipids [367,368]. Inside the CNS, the receptor is mainly expressed in neurons and glial cells, but can also be found in many other tissues outside the brain [369,370]. Six isoforms of RAGE have been identified. The most common variant is integrated into the membrane (mRAGE), while the other five soluble isoforms lack the transmembrane region (sRAGE) and are thus either cytosolic or secreted into the extracellular space [371]. Moreover, RAGE can be cleaved by ADAM10 and ADAM17, which are also involved in APP processing [18]. RAGE was first associated with AD due to its ability to bind oligomeric Aβ1-42, which triggers pro-inflammatory responses in microglia [46,372]. RAGE is involved in Aβ trafficking and clearance through the blood–brain barrier [373] and is thought to modulate secretase activity in neurons, thereby influencing APP processing and Aβ production [374]. Furthermore, RAGE may act as a co-receptor for FPRs that enhances the uptake of Aβ1-42 in microglia [46]. More recently, a possible contribution to AD through interaction with Amylin (also known as islet amyloid polypeptide) was proposed [375]. In microglia of APP/PS1, mice expression of human RAGE exacerbates the production of pro-inflammatory cytokines such as Il1β and TNFα, increases the production of Aβ1-42 and Aβ1-40, and promotes cognitive decline [376]. In contrast, RAGE-deficient APP/PS1 mice showed a decreased concentration of Aβ1-42 and Aβ1-40 and improved spatial memory [374].
FPRs are a small family of G-protein-coupled receptors comprising three receptor subtypes (FPR1, FPR2, and FPR3) in humans [377,378] that are expressed in the innate immune system and a number other cell types, including microglia [379]. FPRs mainly detect formyl methionine-containing peptides that occur in bacteria as PAMPs released during bacterial infections and in mitochondria [380,381], where they are released as DAMPs by damaged cells [381]. They also bind some neuropathological peptides such as Aβ1-42 [231,266] and prion protein fragment PrP106-126 [382]. In particular, FPR2 has been implicated in AD through its interactions with Aβ1-42. In microglia, FPR2 mediates uptake of Aβ1-42 [227,229,383] and triggers the inflammatory response against it. In inflamed brain areas of AD patients, FPR2 is upregulated in reactive glia [231]. Interestingly, microglia from APP/PS1 mice show an increase in the murine FPR1 and FPR2 [46], and treatment of APP/PS1 mice with a competitive FPR inhibitor ameliorated cognitive impairment, reduced plaque load, and lessened microglial reactivity [384]. The role of the FPR subtypes in the response to non-canonical Aβ species is not well examined. However, our recent study provided the first in vitro evidence that FPRs can detect N-terminally abridged Aβ species, and that FPR1 and FPR3 may also be involved in detection of Aβ1-42 [113].
MARCO is a Class A scavenger receptor that is mainly expressed in macrophages and glial cells. It mainly binds poly-anionic ligands such as low-density lipoproteins and environmental particles such as TiO2 and Fe2O3 [385], but has also been shown to capture bacterial PAMPS such as lipopolysaccharides and oxidized lipoproteins, and even helps to engulf whole bacteria in macrophages [386,387,388]. Several studies suggest that MARCO may also bind both fibrillary and non-fibrillary Aβ, either directly or through interactions with other receptors [322]. MARCO modulates intracellular activation of NLRP3 and limits extracellular detection by in glial cells [389]. Furthermore, MARCO can form complexes with FPR2 that facilitate uptake of Aβ1-42 and may influence FPR signaling [319].
Chemokine-like receptor 1 (CMKLR1) is a G-protein-coupled receptor that is mainly expressed in white adipose tissue and immune cells such as microglia, macrophages, and dendritic cells [390]. CMKLR1 interacts with endogenous ligands such as the adipokine Chemerin, and helps to modulate metabolic processes and cell proliferation, especially during glucose processing, adipogenesis, and angiogenesis [390]. In immune cells, CMKLR1 induces chemotaxis and pro-inflammatory cascades, but is also involved in anti-inflammatory signaling through interactions with pro-resolving ligands such as Resolvin E1 [391]. In brain tissue of AD patients, CMKLR1 is upregulated and co-localizes with Aβ1-42 [318]. Furthermore, Aβ1-42 binds to the receptor in an in vitro expression system [318]. In primary mouse microglia and cell lines, interactions between CMKLR1 and Aβ1-42 induce cell migration, internalization of Aβ1-42, and MAPK-dependent inflammatory activation [318]. Knockout of CMKLR1 in APP/PS1 mice leads to increased Aβ deposition but also mortality and cognitive impairment [392]. CMKLR1-deficient mice and wild type mice that were treated with a CMKLR1 inhibitor were more resistant against a chemically induction of Tau hyperphosphorylation [392].
Nucleolin is a phosphoprotein mainly distributed in the nucleolus of many cell types, and is also expressed at the cell surface of macrophages and microglia [330,393]. It can bind DNA, RNA, and amyloid-like proteins [393,394]. Inside the nucleus it has been implicated in regulating DNA and RNA metabolism, chromatin structure, and ribosome assembly [394]. At the cell surface it has been implicated in the cellular-entry of various viruses such as human immunodeficiency virus (HIV) and respiratory syncytial virus (RSV). In primary rat microglia, Nucleolin also robustly recognizes monomeric and fibrillary Aβ1-42 and mediates its phagocytosis, but shows only weak binding to Aβ1-40 [330]. Furthermore, Nucleolin interacts with APP mRNA [395], which may result in modulation and increase APP and Aβ production [396,397].
TRPM2 is a non-selective calcium-permeable cation channel [398]. Activation of TRPM2 is thought to contribute to warmth-sensing in neurons [399] and to the regulation of cytokine secretion in immune cells [400]. The ion channel does not directly interact with classical ligands but instead senses mediators of oxidative stress such as ROS and excessive nitric oxide [401]. In microglia, detection of Aβ1-42 by other receptors leads to the induction of oxidative stress, which in turn activates TRPM2 and induces rapid calcium influx, which triggers cytokine production through the NLRP3 inflammasome [402]. This may lead to a self-renewal of TRPM2 activation since cytokine release can trigger further production of ROS, thus propagating a vicious cycle of pro-inflammatory cascades. Knockout or pharmacological inhibition of TRPM2 in primary mouse microglia attenuates microglial activation by Aβ1-42 and inhibits production of TNFα [403]. Next, TRPM2-deficient APP/PS1 mice show improved spatial memory and reduced microglial reactivity, but no change in plaque load [404].
CD33 is a cell-surface receptor in myeloid cells that is exclusively expressed in microglia within the CNS [405,406]. Through interactions with sialic acids, CD33 can bind glycans and glycolipids. Of note, due to its short extracellular domain, it is assumed that CD33 mainly interacts with ligands on the surface of the cell [406]. In microglia, CD33 does not bind Aβ1-42 but is thought to interact with other receptors—especially with TREM2—to modulate their signaling and inflammatory responses [407]. Several GWASs have identified CD33 as a high-risk factor for the development of AD (see Table 2). Expression of CD33 is increased in microglia of AD patients and is positively correlated with insoluble Aβ1-42 levels and plaque burden [408]. Deletion of the CD33 gene in microglia derived from human-induced pluripotent stem cells or in primary mouse microglia improves uptake of Aβ1-42 [407,408] and leads to increased oxidative burst and production of pro-inflammatory factors [407]. In accordance with these in vitro data, CD33-deficient APP/PS1 mice show reduced brain levels of insoluble Aβ1-42 as well as amyloid plaque burden [408]. Injection of a viral system encoding microRNA against CD33 in APP/PS1 mice resulted in reduced expression of CD33 and decreased cerebral levels of soluble Aβ1-42 and Aβ1-40 [409].
The family of FcRs comprises several receptors that are expressed in microglia and many other immune cells, and can bind the Fc-region of different antibody subtypes. Interestingly, plaque-invading microglia show a strong expression of FcRs [410]. In several animal studies, immunization against Aβ1-42 led to a significant reduction in plaque load [411,412] and removal of Aβ1-42 from CSF [413,414]. A possible explanation is that autoantibodies against Aβ1-42 bind aggregates in senile plaques that are then degraded through FcR-mediated clearance by microglia. However, other studies suggest that these effects are independent from FcRs [415]. During the progression of AD, the blood–brain barrier is weakened, which may allow peripheral cells and molecules to enter the brain [411,412]. This permits the entry of antibodies that can bind to Aβ aggregates. A subsequent binding of this complex to FcRs on microglia may help to clear these pathological peptides from affected brain areas [413,414]
Genome-wide association studies (GWASs) and genetic linkage studies associated with late onset AD (reviewed, e.g., in [416,417]) revealed the importance of genes with a regulatory role for microglia and immune responses. In particular, receptors that are expressed on microglia, such as TREM2 and CD33, have been identified by these approaches. This underpins the key role of microglia in perception of Aβ, the regulation of the brain’s inflammatory response to these deleterious peptides, and the removal of the resulting aggregates. For many other putative microglial Aβ receptors (for a summary see Table 2), such as MARCO, identification of AD-related polymorphisms has not yet been carried out. With regard to “binding to Aβ” of these putative receptors, it has to be taken into account that, in most cases, the direct physical binding has not been proven yet. In most cases, binding has been assumed due to pharmacological or genetic interventions in cells or animals. Exemptions are CD14 or TREM2, where immunoprecipitations were conducted [307,340].
Taken together, these observations argue that a delicate balance of differently modified or processed Aβ variants may finally orchestrate immune responses via a number of different microglial receptors that show a differential presence in different microglia subtypes. A detailed analysis regarding Aβ variant-reactivity or potential selectivity of most receptors is missing to date. Moreover, knowledge on the effect of identified polymorphisms of the receptors regarding recognition of modified/abridged Aβ peptides is even scarcer. Another layer of complexity is added by the fact that some mutations may alter the proteolytic processing, either indirectly through microglial receptor signaling or directly through SNPs in the processing proteases that may affect the cleavage pattern, processivity, and shedding. TREM2 SNPs are a good example of such complex mutual effects. They have been shown to accumulate in cell protrusions in close proximity to Aβ plaques [346] that can be cleaved by the metalloproteases ADAM10/17 C-terminal to histidine 157 [17,418,419,420]. The amount of soluble (s) TREM2 is increased in the CSF of AD patients and seems to sustain microglial viability, and to trigger inflammatory responses by the Akt–GSK3β–β-catenin and NF-κB pathway in vitro [421,422]. Consequently, its beneficial capacity lowers ApoE4-related risk of cognitive decline [423]. Carriers of the R47H TREM2 polymorphism show higher levels of sTREM2 in CSF than non-carriers [347]. The p.H157Y variant that was identified in the Han Chinese population within the stalk region of the receptor enhances shedding of TREM2 [419,424]. Other variants, such as the TREM2 p.T66M mutation, led to reduced secretion of a soluble partition of the receptor [347,350]. While the amount of R47H TREM2 on the cell surface remained stable despite elevated secretion of sTREM2 [350], uncleaved plasma-membrane tethered p.H157Y TREM2 was found to be severely reduced [419]. An association of the two SNPs within the human α-secretase ADAM10 promoter, rs514049A/C and rs653765C/T, with AD could not be clearly demonstrated [425,426]. However, the latter was found to be correlated with lowered ADAM10 mRNA levels in PBMCs of CC versus CT/TT carriers and with a lowered level of soluble (s) RAGE in plasma [426]. RAGE is also expressed on microglia [427] and processed by ADAM10 [428]; therefore, the ADAM10 promoter polymorphisms may also play a role in related inflammatory processes in the brain, even if this has not been investigated yet. Moreover, the RAGE polymorphism G82S itself influences shedding as a significant association between G82S genotypes and sRAGE plasma concentrations in samples from non-diabetic/non-obese Koreans [429].
Membrane topology and architecture of the putative or identified Aβ receptors on microglia are quite diverse (see Figure 1), and the intense phagocytic activity of these cells, in general, requires a dynamic membrane architecture. Therefore, polymorphisms in the receptor-encoding genes themselves or in genes of the interaction partners may not only orchestrate Aβ perception. Lipidomic analyses of microglial CD11b-positive small extracellular vesicles from the cryopreserved parietal cortex of a restricted number of patients and controls indicated not only increased levels of TREM2, but also a proinflammatory lipid profile in AD (e.g., increase in the most abundant monohexosylceramide d18:1/24:1 [430]). Moreover, an increase in cholesterol was observed that might influence fluidity of the membrane, in addition to other receptor-modifying pathways such as synthesis of ligands. Sequestration of cholesterol in astrocytes, for example, affected APP processing and accumulation of Aβ peptides [431], and serum starvation induced shedding of BACE1, which could be further aggravated by cholesterol efflux mediated by methyl β cyclodextran [432]. TREM2 itself has been shown to act as a regulator of brain cholesterol metabolism, and TREM2-deficient microglia fail to degrade myelin-derived cholesterol [433]. How its polymorphisms or binding of different Aβ variants might interfere with this function has not been addressed to the best of our knowledge.
In sum, this indicates that activation of microglia by Aβ variants can be affected not only directly by genetic variants of the receptors, but also by their impact on receptor processing or on other functions of the receptors that subsequently reflect on Aβ-driven signaling pathways. A deeper understanding of relevance of the single observed variants for the distinct receptor functions on the microglial surface therefore has to be the focus of research in the future to estimate their potential as therapeutic targets or their relevance in individualized medicine.

6. Secondary Structure and Oligomerization Critically Affect Aβ Neurotoxicity

Different aggregate forms of Aβ that lead to plagues have long been suspected to be the major culprit of AD. However, there is clear evidence indicating that the formation of insoluble aggregates alone is insufficient to trigger typical neuroinflammation and neurodegeneration. First, the formation of insoluble Aβ plaques also occurs during normal aging in many people who do not suffer from AD [434,435]. Second, some patients display typical AD symptoms but show only low plaque formation [436]. Third, transgenic APP mice already show a pronounced plaque formation before the onset of neuroinflammatory events [437]. Finally, therapeutic approaches for plaque removal did not ameliorate cognitive symptoms in human patients [9]. This indicates that at least a second factor has to contribute. Increasing evidence suggests that the formation of soluble oligomers and/or differences in the secondary structure might be the second culprit that triggers the initial inflammatory events. For example, soluble Aβ1-42 oligomers are highly neurotoxic and can induce strong pro-inflammatory activity in glial cells [438,439]. Next, the Osaka mutation, a familiar form of AD that causes a loss of glutamate at position 22 of Aβ, leads to Aβ peptides with enhanced soluble oligomer formation capability but which are unable to form fibrils, and can therefore not be deposited as insoluble plaques (reviewed in [440]). Studies that injected soluble Aβ1-42 oligomers into healthy rodents’ brains observed strong neurodegenerative effects, including aberrant neuroinflammation, synaptic disruption, neuronal death, and cognitive deficits [441,442,443]. For example, Aβ*56, a distinct species of Aβ oligomers with a molecular weight of 56 kDa, was proposed as a promising candidate to explain the genesis of AD, since the injection of such aggregates could induce strong cognitive decline in rat models [444]. However, recent investigations have questioned the validity of these studies [445]. In addition, other studies could not detect Aβ*56 in tissue [439,446] or CSF samples [438,446] of human AD patients or AD model mice. Next, a number of other studies reported no or only mild pathological effects of Aβ oligomer injection into the brain of animals [447,448]. The interpretation of the divergent results is still problematic. However, technical issues cannot be excluded because these studies used a plethora of different protocols that varied strongly in terms of buffer composition, pH value, aggregation kinetics, Aβ concentration, temperature, aggregation time, and agitation. Numerous studies have demonstrated that even small deviations in the environmental conditions lead to drastic differences in secondary structure, the solubility of the resulting aggregates, and their capabilities to elicit biological effects [449,450].
The secondary and tertiary structure of Aβ is an extremely critical factor for its oligomerization and bio-activity (reviewed in [12]). The secondary structure of Aβ1-40 and Aβ1-42 may assume multiple discrete conformations with α-helix or β-sheet conformers, which can undergo rapid changes depending on environmental factors [451,452]. Aβ1-42 possesses two motifs that are capable of forming β-sheets: the highly hydrophobic core motif KLVFFAE (Aβ16-22) [453] and the C-terminal region IIGLMVGGVVIA (Aβ30-41) [451]. β-sheet formation is essential for aggregation, since β-sheet regions of individual peptides self-assemble into cross-linked structures with parallel or antiparallel organization (cross-β patterns) that self-assemble interlinked fibrils and protofibrils [12,454,455]. This self-assembly from monomers in solution is first auto-catalyzed by the formation of nuclei, small aggregates with especially high thermodynamic energy states. These energy levels lead to faster association than dissociation rates of monomers and small oligomers to the nucleus and thus drive further aggregation [456]. This process is enhanced through secondary nucleation where other fibrils or protofibrils in the solution associate with the initial nucleus and accelerate its growth. The nucleation and growth phase finally ends in fibril formation since the fibrillary state is thermodynamically the most stable form of any peptide assembly [456]. In particular, the inflexible hydrophobic C-terminus of Aβ is thought to initiate the transformation from α-helical to β-sheet structure that is needed for the formation of nuclei [12,451]. Accordingly, C-terminal truncation decreases the hydrophobic region and increases the structural flexibility of Aβ, which in turn decreases its propensity to form β-sheets. This may explain why Aβ species with longer C-termini tend to be more amyloidogenic then shorter variants.
Of note, many soluble Aβ variants and their oligomers are present as metastable species with distinct thermodynamic energy levels [457,458]. Some studies have proposed the existence of a α-pleated sheet conformation in a subset of these soluble Aβ oligomers that might set them apart from non-toxic fibrils and are not generated by typical nucleation [459,460,461]. The α-pleated sheet conformation is thought to be structurally similar to β-sheets but may exhibit a distinct biological effect. However, the biological relevance of α-pleated sheet conformations is still debated. A recent study by Shea and colleagues utilized probes designed to catch α-pleated sheet peptides in a transgenic Aβ Caenorhabditis elegans model and in transgenic APP mice, which resulted in a reduction in soluble Aβ oligomers in both models [460].
Taken together, these findings highlight the high importance of keeping track of Aβ’s secondary structure during the investigation of its biological effects. This is especially necessary in studies performed with synthetic Aβ peptides. In general, synthesis processes are not performed under physiological conditions and, therefore, can produce peptides with altered conformations [462]. During synthesis, Aβ peptides already aggregate; thus, there is little control over the Aβ species that comprise the final end product [462]. Next, most peptide synthesis processes depend on counter ions such as trifluoracetate or hydrochloride, which are bound to the final synthesis product and can highly influence the solubility, nucleation, and aggregation kinetic of generated Aβ peptides [463,464]. Thus, the use of Aβ peptides derived from different companies and synthesis processes can produce largely different biological effects [113,465], which may additionally vary from batch to batch [456,465]. Various methods have been developed to minimize these aggregation artefacts, e.g., dissolving synthetic Aβ in harsh solvents such as DMSO, HFIP, or NH4OH [450,466,467]. However, several studies provided evidence that these treatments can compromise the secondary structure of Aβ [468,469].
Taken together, these findings highlight the need for standardization protocols that improve the comparability and reproducibility of Aβ research. All in vitro experiments should be performed with Aβ obtained from at least two independent sources to validate the observed biological effects and to avoid potential artifacts [113,456,465]. Inexpensive methods such as Thioflavin T aggregation assays or analysis via SDS-PAGE should be used to give some insight regarding the aggregation state of Aβ. Next, all experimental conditions have to be described very carefully because the exact assay buffer composition, pH value, incubation time, assay temperature, and Aβ storage conditions can all influence the outcome of experiments. Finally, it is essential to also report negative or non-conclusive data obtained with different Aβ variants and manufacturers, since these can still provide valuable information for other researchers and may help to elucidate the bigger picture of the still mysterious Aβ.

7. Concluding Remarks

In summary, the currently available data clearly argue for a significant contribution of non-canonical Aβ variants to the pathogenesis and progression of AD. However, there is still an amazingly large lack of knowledge on the precise contribution of most of these variants. Next, it seems that the influence of gene polymorphisms regarding the recognition of modified or abridged Aβ peptides is even scarcer. Amazingly, only 20 polymorphisms of Aβ interaction partners have so far been clearly associated with AD. Given that for TREM2 alone nearly 200 single nucleotide exchanges have been reported, and that at least 16 other interaction partners exist, presumably thousands of additional SNPs still await a careful analysis of their impact on receptor interaction, proteolytic processing, and shedding. Next, there is clear evidence that microglia, astrocytes, and oligodendrocytes are key players in the production and release of non-canonical Aβ variants. Therefore, the interplay of these cell types with neurons has to be better characterized. Finally, there is a clear need for standardized protocols in in vitro studies and more information on the Aβ peptide composition in in vivo studies. The current literature already provides ample evidence that chemical modifications, extracellular environment, and precise quantities of different amyloid species have a strong influence on the aggregation and bio-activity of Aβ variants. In a recent study, we even showed that supposedly identical Aβ peptides are highly subjectable to solvent- and manufacturer-dependent effects [113]. For the sake of reproducibility, it is thus of uttermost importance to work with more than one peptide in all in vitro experiments, to very precisely report all experimental conditions, and to include data on the secondary structure, 3D conformation, and aggregation kinetics [113,456,465]. In vivo studies urgently need to gather more information on the precise Aβ variant composition, the existing oligomers, and their 3D structure. This will help to identify the key variants and their structural requirements for a given physiological effect.

Author Contributions

Conceptualization, L.B., S.E., K.E. and B.B.; writing—original draft preparation, L.B., S.E., K.E. and B.B.; writing—review and editing, L.B., S.E., K.E. and B.B.; visualization, L.B., S.E., K.E. and B.B.; funding acquisition, K.E. and B.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by funding from the Ministerium für Wissenschaft und Gesundheit (MWG), Rheinland Pfalz, NeurodegX Forschungskolleg and MultiSense Forschungskolleg.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

All figures were created using BioRender.com and exported under a paid license. The 3D models in Figure 3A were rendered using the software Virtual Molecular Dynamics (VMD) [470].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akiyama, H.; Arai, T.; Kondo, H.; Tanno, E.; Haga, C.; Ikeda, K. Cell mediators of inflammation in the Alzheimer disease brain. Alzheimer Dis. Assoc. Disord. 2000, 14 (Suppl. 1), S47–S53. [Google Scholar] [CrossRef] [PubMed]
  2. Heneka, M.T.; Kummer, M.P.; Latz, E. Innate immune activation in neurodegenerative disease. Nat. Rev. Immunol. 2014, 14, 463–477. [Google Scholar] [CrossRef] [PubMed]
  3. Leng, F.; Edison, P. Neuroinflammation and microglial activation in Alzheimer disease: Where do we go from here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef] [PubMed]
  4. Rogers, J.; Luber-Narod, J.; Styren, S.D.; Civin, W.H. Expression of immune system-associated antigens by cells of the human central nervous system: Relationship to the pathology of Alzheimer’s disease. Neurobiol. Aging 1988, 9, 339–349. [Google Scholar] [CrossRef]
  5. McGeer, P.L.; Itagaki, S.; Tago, H.; McGeer, E.G. Occurrence of HLA-DR reactive microglia in Alzheimer’s disease. Ann. N. Y. Acad. Sci. 1988, 540, 319–323. [Google Scholar] [CrossRef] [PubMed]
  6. Itagaki, S.; McGeer, P.L.; Akiyama, H.; Zhu, S.; Selkoe, D. Relationship of microglia and astrocytes to amyloid deposits of Alzheimer disease. J. Neuroimmunol. 1989, 24, 173–182. [Google Scholar] [CrossRef]
  7. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  8. Masters, C.L.; Multhaup, G.; Simms, G.; Pottgiesser, J.; Martins, R.N.; Beyreuther, K. Neuronal origin of a cerebral amyloid: Neurofibrillary tangles of Alzheimer’s disease contain the same protein as the amyloid of plaque cores and blood vessels. EMBO J. 1985, 4, 2757–2763. [Google Scholar] [CrossRef]
  9. Haass, C.; Selkoe, D. If amyloid drives Alzheimer disease, why have anti-amyloid therapies not yet slowed cognitive decline? PLoS Biol. 2022, 20, e3001694. [Google Scholar] [CrossRef]
  10. Steiner, H.; Fukumori, A.; Tagami, S.; Okochi, M. Making the final cut: Pathogenic amyloid-beta peptide generation by gamma-secretase. Cell. Stress 2018, 2, 292–310. [Google Scholar] [CrossRef]
  11. Andreasson, U.; Portelius, E.; Andersson, M.E.; Blennow, K.; Zetterberg, H. Aspects of beta-amyloid as a biomarker for Alzheimer’s disease. Biomark. Med. 2007, 1, 59–78. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, G.F.; Xu, T.H.; Yan, Y.; Zhou, Y.R.; Jiang, Y.; Melcher, K.; Xu, H.E. Amyloid beta: Structure, biology and structure-based therapeutic development. Acta. Pharm. Sin. 2017, 38, 1205–1235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kummer, M.P.; Heneka, M.T. Truncated and modified amyloid-beta species. Alzheimers Res. 2014, 6, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Doens, D.; Fernández, P.L. Microglia receptors and their implications in the response to amyloid β for Alzheimer’s disease pathogenesis. J. Neuroinflammation 2014, 11, 48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Busch, L.; Vieten, S.; Brödel, S.; Endres, K.; Bufe, B. Emerging contributions of formyl peptide receptors to neurodegenerative diseases. Biol. Chem. 2022, 403, 27–41. [Google Scholar] [CrossRef]
  16. Gospodarska, E.; Kupniewska-Kozak, A.; Goch, G.; Dadlez, M. Binding studies of truncated variants of the Abeta peptide to the V-domain of the RAGE receptor reveal Abeta residues responsible for binding. Biochim. Biophys. Acta 2011, 1814, 592–609. [Google Scholar] [CrossRef]
  17. Wunderlich, P.; Glebov, K.; Kemmerling, N.; Tien, N.T.; Neumann, H.; Walter, J. Sequential proteolytic processing of the triggering receptor expressed on myeloid cells-2 (TREM2) protein by ectodomain shedding and gamma-secretase-dependent intramembranous cleavage. J. Biol. Chem. 2013, 288, 33027–33036. [Google Scholar] [CrossRef] [Green Version]
  18. Raucci, A.; Cugusi, S.; Antonelli, A.; Barabino, S.M.; Monti, L.; Bierhaus, A.; Reiss, K.; Saftig, P.; Bianchi, M.E. A soluble form of the receptor for advanced glycation endproducts (RAGE) is produced by proteolytic cleavage of the membrane-bound form by the sheddase a disintegrin and metalloprotease 10 (ADAM10). FASEB J. 2008, 22, 3716–3727. [Google Scholar] [CrossRef]
  19. Okello, A.; Edison, P.; Archer, H.A.; Turkheimer, F.E.; Kennedy, J.; Bullock, R.; Walker, Z.; Kennedy, A.; Fox, N.; Rossor, M.; et al. Microglial activation and amyloid deposition in mild cognitive impairment: A PET study. Neurology 2009, 72, 56–62. [Google Scholar] [CrossRef] [Green Version]
  20. Hamelin, L.; Lagarde, J.; Dorothee, G.; Leroy, C.; Labit, M.; Comley, R.A.; de Souza, L.C.; Corne, H.; Dauphinot, L.; Bertoux, M.; et al. Early and protective microglial activation in Alzheimer’s disease: A prospective study using 18F-DPA-714 PET imaging. Brain 2016, 139, 1252–1264. [Google Scholar] [CrossRef]
  21. Femminella, G.D.; Dani, M.; Wood, M.; Fan, Z.; Calsolaro, V.; Atkinson, R.; Edginton, T.; Hinz, R.; Brooks, D.J.; Edison, P. Microglial activation in early Alzheimer trajectory is associated with higher gray matter volume. Neurology 2019, 92, e1331–e1343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Nimmerjahn, A.; Kirchhoff, F.; Helmchen, F. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 2005, 308, 1314–1318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Rodríguez-Gómez, J.A.; Kavanagh, E.; Engskog-Vlachos, P.; Engskog, M.K.R.; Herrera, A.J.; Espinosa-Oliva, A.M.; Joseph, B.; Hajji, N.; Venero, J.L.; Burguillos, M.A. Microglia: Agents of the CNS Pro-Inflammatory Response. Cells 2020, 9, 1717. [Google Scholar] [CrossRef] [PubMed]
  24. Li, Q.; Barres, B.A. Microglia and macrophages in brain homeostasis and disease. Nat. Rev. Immunol. 2018, 18, 225–242. [Google Scholar] [CrossRef]
  25. Tremblay, M.E.; Stevens, B.; Sierra, A.; Wake, H.; Bessis, A.; Nimmerjahn, A. The Role of Microglia in the Healthy Brain. J. Neurosci. 2011, 31, 16064–16069. [Google Scholar] [CrossRef] [Green Version]
  26. Brionne, T.C.; Tesseur, I.; Masliah, E.; Wyss-Coray, T. Loss of TGF-β1 Leads to Increased Neuronal Cell Death and Microgliosis in Mouse Brain. Neuron 2003, 40, 1133–1145. [Google Scholar] [CrossRef] [Green Version]
  27. Labandeira-Garcia, J.L.; Costa-Besada, M.A.; Labandeira, C.M.; Villar-Cheda, B.; Rodriguez-Perez, A.I. Insulin-Like Growth Factor-1 and Neuroinflammation. Front. Aging Neurosci. 2017, 9, 365. [Google Scholar] [CrossRef] [Green Version]
  28. Wake, H.; Moorhouse, A.J.; Jinno, S.; Kohsaka, S.; Nabekura, J. Resting Microglia Directly Monitor the Functional State of Synapses In Vivo and Determine the Fate of Ischemic Terminals. J. Neurosci. 2009, 29, 3974–3980. [Google Scholar] [CrossRef] [Green Version]
  29. Paolicelli, R.C.; Bolasco, G.; Pagani, F.; Maggi, L.; Scianni, M.; Panzanelli, P.; Giustetto, M.; Ferreira, T.A.; Guiducci, E.; Dumas, L.; et al. Synaptic pruning by microglia is necessary for normal brain development. Science 2011, 333, 1456–1458. [Google Scholar] [CrossRef] [Green Version]
  30. Harley, S.B.R.; Willis, E.F.; Shaikh, S.N.; Blackmore, D.G.; Sah, P.; Ruitenberg, M.J.; Bartlett, P.F.; Vukovic, J. Selective Ablation of BDNF from Microglia Reveals Novel Roles in Self-Renewal and Hippocampal Neurogenesis. J. Neurosci. 2021, 41, 4172–4186. [Google Scholar] [CrossRef]
  31. Parkhurst, C.N.; Yang, G.; Ninan, I.; Savas, J.N.; Yates, J.R., 3rd; Lafaille, J.J.; Hempstead, B.L.; Littman, D.R.; Gan, W.B. Microglia promote learning-dependent synapse formation through brain-derived neurotrophic factor. Cell 2013, 155, 1596–1609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Friedman, B.A.; Srinivasan, K.; Ayalon, G.; Meilandt, W.J.; Lin, H.; Huntley, M.A.; Cao, Y.; Lee, S.H.; Haddick, P.C.G.; Ngu, H.; et al. Diverse Brain Myeloid Expression Profiles Reveal Distinct Microglial Activation States and Aspects of Alzheimer’s Disease Not Evident in Mouse Models. Cell. Rep. 2018, 22, 832–847. [Google Scholar] [CrossRef] [Green Version]
  33. Masuda, T.; Sankowski, R.; Staszewski, O.; Prinz, M. Microglia Heterogeneity in the Single-Cell Era. Cell. Rep. 2020, 30, 1271–1281. [Google Scholar] [CrossRef] [PubMed]
  34. Keren-Shaul, H.; Spinrad, A.; Weiner, A.; Matcovitch-Natan, O.; Dvir-Szternfeld, R.; Ulland, T.K.; David, E.; Baruch, K.; Lara-Astaiso, D.; Toth, B.; et al. A Unique Microglia Type Associated with Restricting Development of Alzheimer’s Disease. Cell 2017, 169, 1276–1290.e17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Deczkowska, A.; Keren-Shaul, H.; Weiner, A.; Colonna, M.; Schwartz, M.; Amit, I. Disease-Associated Microglia: A Universal Immune Sensor of Neurodegeneration. Cell 2018, 173, 1073–1081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta. Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Attwell, D.; Buchan, A.M.; Charpak, S.; Lauritzen, M.; Macvicar, B.A.; Newman, E.A. Glial and neuronal control of brain blood flow. Nature 2010, 468, 232–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Abbott, N.J.; Rönnbäck, L.; Hansson, E. Astrocyte–endothelial interactions at the blood–brain barrier. Nat. Rev. Neurosci. 2006, 7, 41–53. [Google Scholar] [CrossRef]
  39. Heithoff, B.P.; George, K.K.; Phares, A.N.; Zuidhoek, I.A.; Munoz-Ballester, C.; Robel, S. Astrocytes are necessary for blood–brain barrier maintenance in the adult mouse brain. Glia 2021, 69, 436–472. [Google Scholar] [CrossRef]
  40. Iliff, J.J.; Wang, M.; Liao, Y.; Plogg, B.A.; Peng, W.; Gundersen, G.A.; Benveniste, H.; Vates, G.E.; Deane, R.; Goldman, S.A.; et al. A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid β. Sci. Transl. Med. 2012, 4, 147ra111. [Google Scholar] [CrossRef]
  41. Jessen, N.A.; Munk, A.S.; Lundgaard, I.; Nedergaard, M. The Glymphatic System: A Beginner’s Guide. Neurochem. Res. 2015, 40, 2583–2599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends. Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef] [PubMed]
  43. Liddelow, S.A.; Barres, B.A. Reactive Astrocytes: Production, Function, and Therapeutic Potential. Immunity 2017, 46, 957–967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Konishi, H.; Okamoto, T.; Hara, Y.; Komine, O.; Tamada, H.; Maeda, M.; Osako, F.; Kobayashi, M.; Nishiyama, A.; Kataoka, Y.; et al. Astrocytic phagocytosis is a compensatory mechanism for microglial dysfunction. Embo. J. 2020, 39, e104464. [Google Scholar] [CrossRef]
  45. Gorina, R.; Font-Nieves, M.; Márquez-Kisinousky, L.; Santalucia, T.; Planas, A.M. Astrocyte TLR4 activation induces a proinflammatory environment through the interplay between MyD88-dependent NFκB signaling, MAPK, and Jak1/Stat1 pathways. Glia 2011, 59, 242–255. [Google Scholar] [CrossRef] [PubMed]
  46. Slowik, A.; Merres, J.; Elfgen, A.; Jansen, S.; Mohr, F.; Wruck, C.J.; Pufe, T.; Brandenburg, L.-O. Involvement of formyl peptide receptors in receptor for advanced glycation end products (RAGE) - and amyloid beta 1-42-induced signal transduction in glial cells. Mol. Neurodegener. 2012, 7, 55. [Google Scholar] [CrossRef] [Green Version]
  47. Mitew, S.; Kirkcaldie, M.T.; Dickson, T.C.; Vickers, J.C. Altered synapses and gliotransmission in Alzheimer’s disease and AD model mice. Neurobiol. Aging 2013, 34, 2341–2351. [Google Scholar] [CrossRef]
  48. Yamamoto, N.; Ishikuro, R.; Tanida, M.; Suzuki, K.; Ikeda-Matsuo, Y.; Sobue, K. Insulin-signaling Pathway Regulates the Degradation of Amyloid β-protein via Astrocytes. Neuroscience 2018, 385, 227–236. [Google Scholar] [CrossRef]
  49. Sbai, O.; Ould-Yahoui, A.; Ferhat, L.; Gueye, Y.; Bernard, A.; Charrat, E.; Mehanna, A.; Risso, J.-J.; Chauvin, J.-P.; Fenouillet, E.; et al. Differential vesicular distribution and trafficking of MMP-2, MMP-9, and their inhibitors in astrocytes. Glia 2010, 58, 344–366. [Google Scholar] [CrossRef]
  50. González-Reyes, R.E.; Nava-Mesa, M.O.; Vargas-Sánchez, K.; Ariza-Salamanca, D.; Mora-Muñoz, L. Involvement of Astrocytes in Alzheimer’s Disease from a Neuroinflammatory and Oxidative Stress Perspective. Front. Mol. Neurosci. 2017, 10, 427. [Google Scholar] [CrossRef]
  51. Spampinato, S.F.; Merlo, S.; Fagone, E.; Fruciano, M.; Sano, Y.; Kanda, T.; Sortino, M.A. Reciprocal Interplay Between Astrocytes and CD4+ Cells Affects Blood-Brain Barrier and Neuronal Function in Response to β Amyloid. Front. Mol. Neurosci. 2020, 13. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, D.; Chen, F.; Han, Z.; Yin, Z.; Ge, X.; Lei, P. Relationship Between Amyloid-β Deposition and Blood–Brain Barrier Dysfunction in Alzheimer’s Disease. Front. Cell. Neurosci. 2021, 15. [Google Scholar] [CrossRef] [PubMed]
  53. Spampinato, S.F.; Merlo, S.; Sano, Y.; Kanda, T.; Sortino, M.A. Astrocytes contribute to Aβ-induced blood–brain barrier damage through activation of endothelial MMP9. J. Neurochem. 2017, 142, 464–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Bradl, M.; Lassmann, H. Oligodendrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 37–53. [Google Scholar] [CrossRef] [Green Version]
  55. Nave, K.-A.; Werner, H.B. Ensheathment and Myelination of Axons: Evolution of Glial Functions. Annu. Rev. Neurosci. 2021, 44, 197–219. [Google Scholar] [CrossRef]
  56. Church, J.S.; Kigerl, K.A.; Lerch, J.K.; Popovich, P.G.; McTigue, D.M. TLR4 Deficiency Impairs Oligodendrocyte Formation in the Injured Spinal Cord. J. Neurosci. 2016, 36, 6352–6364. [Google Scholar] [CrossRef] [Green Version]
  57. Grasselli, C.; Ferrari, D.; Zalfa, C.; Soncini, M.; Mazzoccoli, G.; Facchini, F.A.; Marongiu, L.; Granucci, F.; Copetti, M.; Vescovi, A.L.; et al. Toll-like receptor 4 modulation influences human neural stem cell proliferation and differentiation. Cell. Death Dis.. 2018, 9, 280. [Google Scholar] [CrossRef] [Green Version]
  58. Qin, J.; Goswami, R.; Dawson, S.; Dawson, G. Expression of the receptor for advanced glycation end products in oligodendrocytes in response to oxidative stress. J. Neurosci. Res. 2008, 86, 2414–2422. [Google Scholar] [CrossRef] [Green Version]
  59. Cain, A.; Taga, M.; McCabe, C.; Green, G.; Hekselman, I.; White, C.C.; Lee, D.I.; Gaur, P.; Rozenblatt-Rosen, O.; Zhang, F.; et al. Multi-cellular communities are perturbed in the aging human brain and Alzheimer’s disease. bioRxiv 2022. [Google Scholar] [CrossRef]
  60. Zhou, Y.; Song, W.M.; Andhey, P.S.; Swain, A.; Levy, T.; Miller, K.R.; Poliani, P.L.; Cominelli, M.; Grover, S.; Gilfillan, S.; et al. Human and mouse single-nucleus transcriptomics reveal TREM2-dependent and TREM2-independent cellular responses in Alzheimer’s disease. Nat. Med. 2020, 26, 131–142. [Google Scholar] [CrossRef]
  61. Leng, K.; Li, E.; Eser, R.; Piergies, A.; Sit, R.; Tan, M.; Neff, N.; Li, S.H.; Rodriguez, R.D.; Suemoto, C.K.; et al. Molecular characterization of selectively vulnerable neurons in Alzheimer’s disease. Nat. Neurosci. 2021, 24, 276–287. [Google Scholar] [CrossRef] [PubMed]
  62. Lau, S.-F.; Cao, H.; Fu, A.K.Y.; Ip, N.Y. Single-nucleus transcriptome analysis reveals dysregulation of angiogenic endothelial cells and neuroprotective glia in Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2020, 117, 25800–25809. [Google Scholar] [CrossRef] [PubMed]
  63. Mathys, H.; Davila-Velderrain, J.; Peng, Z.; Gao, F.; Mohammadi, S.; Young, J.Z.; Menon, M.; He, L.; Abdurrob, F.; Jiang, X.; et al. Single-cell transcriptomic analysis of Alzheimer’s disease. Nature 2019, 570, 332–337. [Google Scholar] [CrossRef] [PubMed]
  64. Grubman, A.; Chew, G.; Ouyang, J.F.; Sun, G.; Choo, X.Y.; McLean, C.; Simmons, R.K.; Buckberry, S.; Vargas-Landin, D.B.; Poppe, D.; et al. A single-cell atlas of entorhinal cortex from individuals with Alzheimer’s disease reveals cell-type-specific gene expression regulation. Nat. Neurosci. 2019, 22, 2087–2097. [Google Scholar] [CrossRef] [PubMed]
  65. Pandey, S.; Shen, K.; Lee, S.H.; Shen, Y.A.; Wang, Y.; Otero-García, M.; Kotova, N.; Vito, S.T.; Laufer, B.I.; Newton, D.F.; et al. Disease-associated oligodendrocyte responses across neurodegenerative diseases. Cell. Rep. 2022, 40, 111189. [Google Scholar] [CrossRef]
  66. Caso, F.; Agosta, F.; Mattavelli, D.; Migliaccio, R.; Canu, E.; Magnani, G.; Marcone, A.; Copetti, M.; Falautano, M.; Comi, G.; et al. White Matter Degeneration in Atypical Alzheimer Disease. Radiology 2015, 277, 162–172. [Google Scholar] [CrossRef]
  67. Lee, S.; Viqar, F.; Zimmerman, M.E.; Narkhede, A.; Tosto, G.; Benzinger, T.L.; Marcus, D.S.; Fagan, A.M.; Goate, A.; Fox, N.C.; et al. White matter hyperintensities are a core feature of Alzheimer’s disease: Evidence from the dominantly inherited Alzheimer network. Ann. Neurol. 2016, 79, 929–939. [Google Scholar] [CrossRef] [Green Version]
  68. Bartzokis, G. Alzheimer’s disease as homeostatic responses to age-related myelin breakdown. Neurobiol. Aging 2011, 32, 1341–1371. [Google Scholar] [CrossRef] [Green Version]
  69. Depp, C.; Sun, T.; Sasmita, A.O.; Spieth, L.; Berghoff, S.A.; Steixner-Kumar, A.A.; Subramanian, S.; Möbius, W.; Göbbels, S.; Saher, G.; et al. Ageing-associated myelin dysfunction drives amyloid deposition in mouse models of Alzheimer’s disease. bioRxiv. 2021. [Google Scholar] [CrossRef]
  70. Wildburger, N.C.; Esparza, T.J.; Leduc, R.D.; Fellers, R.T.; Thomas, P.M.; Cairns, N.J.; Kelleher, N.L.; Bateman, R.J.; Brody, D.L. Diversity of Amyloid-beta Proteoforms in the Alzheimer’s Disease Brain. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef]
  71. Abraham, J.D.; Prome, S.; Salvetat, N.; Rubrecht, L.; Cobo, S.; du Paty, E.; Galea, P.; Mathieu-Dupas, E.; Ranaldi, S.; Caillava, C.; et al. Cerebrospinal Abeta11-x and 17-x levels as indicators of mild cognitive impairment and patients’ stratification in Alzheimer’s disease. Transl. Psychiatry 2013, 3, e281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Moro, M.L.; Phillips, A.S.; Gaimster, K.; Paul, C.; Mudher, A.; Nicoll, J.A.R.; Boche, D. Pyroglutamate and Isoaspartate modified Amyloid-Beta in ageing and Alzheimer’s disease. Acta Neuropathol. Commun. 2018, 6, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Weggen, S.; Beher, D. Molecular consequences of amyloid precursor protein and presenilin mutations causing autosomal-dominant Alzheimer’s disease. Alzheimers Res. 2012, 4, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Brinkmalm, G.; Portelius, E.; Öhrfelt, A.; Mattsson, N.; Persson, R.; Gustavsson, M.K.; Vite, C.H.; Gobom, J.; Månsson, J.E.; Nilsson, J.; et al. An online nano-LC-ESI-FTICR-MS method for comprehensive characterization of endogenous fragments from amyloid β and amyloid precursor protein in human and cat cerebrospinal fluid. J. Mass. Spectrom. JMS 2012, 47, 591–603. [Google Scholar] [CrossRef]
  75. Maddalena, A.S.; Papassotiropoulos, A.; Gonzalez-Agosti, C.; Signorell, A.; Hegi, T.; Pasch, T.; Nitsch, R.M.; Hock, C. Cerebrospinal fluid profile of amyloid β peptides in patients with Alzheimer’s disease determined by protein biochip technology. Neurodegener. Dis. 2004, 1, 231–235. [Google Scholar] [CrossRef] [Green Version]
  76. Halim, A.; Brinkmalm, G.; Rüetschi, U.; Westman-Brinkmalm, A.; Portelius, E.; Zetterberg, H.; Blennow, K.; Larson, G.; Nilsson, J. Site-specific characterization of threonine, serine, and tyrosine glycosylations of amyloid precursor protein/amyloid β-peptides in human cerebrospinal fluid. Proc. Natl. Acad. Sci. USA 2011, 108, 11848–11853. [Google Scholar] [CrossRef] [Green Version]
  77. Güntert, A.; Döbeli, H.; Bohrmann, B. High sensitivity analysis of amyloid-beta peptide composition in amyloid deposits from human and PS2APP mouse brain. Neuroscience 2006, 143, 461–475. [Google Scholar] [CrossRef]
  78. Portelius, E.; Bogdanovic, N.; Gustavsson, M.K.; Volkmann, I.; Brinkmalm, G.; Zetterberg, H.; Winblad, B.; Blennow, K. Mass spectrometric characterization of brain amyloid beta isoform signatures in familial and sporadic Alzheimer’s disease. Acta Neuropathol. 2010, 120, 185–193. [Google Scholar] [CrossRef] [Green Version]
  79. Näslund, J.; Schierhorn, A.; Hellman, U.; Lannfelt, L.; Roses, A.D.; Tjernberg, L.O.; Silberring, J.; Gandy, S.E.; Winblad, B.; Greengard, P.; et al. Relative abundance of Alzheimer Aβ amyloid peptide variants in Alzheimer disease and normal aging. Proc. Natl. Acad. Sci. USA 1994, 91, 8378–8382. [Google Scholar] [CrossRef] [Green Version]
  80. Gkanatsiou, E.; Portelius, E.; Toomey, C.E.; Blennow, K.; Zetterberg, H.; Lashley, T.; Brinkmalm, G. A distinct brain beta amyloid signature in cerebral amyloid angiopathy compared to Alzheimer’s disease. Neurosci. Lett. 2019, 701, 125–131. [Google Scholar] [CrossRef]
  81. Gowing, E.; Roher, A.E.; Woods, A.S.; Cotter, R.J.; Chaney, M.; Little, S.P.; Ball, M.J. Chemical characterization of Aβ 17-42 peptide, a component of diffuse amyloid deposits of Alzheimer disease. J. Biol. Chem. 1994, 269, 10987–10990. [Google Scholar] [CrossRef]
  82. Kubo, T.; Kumagae, Y.; Miller, C.A.; Kaneko, I. β-amyloid racemized at the Ser26 residue in the brains of patients with Alzheimer disease: Implications in the pathogenesis of Alzheimer disease. J. Neuropathol. Exp. Neurol. 2003, 62, 248–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Portelius, E.; Lashley, T.; Westerlund, A.; Persson, R.; Fox, N.C.; Blennow, K.; Revesz, T.; Zetterberg, H. Brain amyloid-beta fragment signatures in pathological ageing and alzheimer’s disease by hybrid immunoprecipitation mass spectrometry. Neurodegener. Dis. 2015, 15, 50–57. [Google Scholar] [CrossRef] [PubMed]
  84. Inoue, K.; Hosaka, D.; Mochizuki, N.; Akatsu, H.; Tsutsumiuchi, K.; Hashizume, Y.; Matsukawa, N.; Yamamoto, T.; Toyo’Oka, T. Simultaneous determination of post-translational racemization and isomerization of N -terminal amyloid-β in alzheimer’s brain tissues by covalent chiral derivatized ultraperformance liquid chromatography tandem mass spectrometry. Anal. Chem. 2014, 86, 797–804. [Google Scholar] [CrossRef]
  85. Pekov, S.I.; Ivanov, D.G.; Bugrova, A.E.; Indeykina, M.I.; Zakharova, V.N.; Popov, I.A.; Kononikhin, A.S.; Kozin, S.A.; Makarov, A.A.; Nikolaev, E.N. Evaluation of MALDI-TOF/TOF Mass Spectrometry Approach for Quantitative Determination of Aspartate Residue Isomerization in the Amyloid-β Peptide. J. Am. Soc. Mass. Spectrom. 2019, 30, 1325–1329. [Google Scholar] [CrossRef] [PubMed]
  86. Alzforum. Mutations Database. Available online: https://www.alzforum.org/mutations/search (accessed on 27 October 2022).
  87. Maler, J.M.; Spitzer, P.; Klafki, H.W.; Esselmann, H.; Lewczuk, P.; Kornhuber, J.; Herrmann, M.; Wiltfang, J. Distinct fractional Abeta release patterns in human mononuclear phagocytes. J. Neuroimmunol. 2009, 206, 1–4. [Google Scholar] [CrossRef]
  88. Giau, V.V.; Bagyinszky, E.; Yang, Y.S.; Youn, Y.C.; An, S.S.A.; Kim, S.Y. Genetic analyses of early-onset Alzheimer’s disease using next generation sequencing. Sci. Rep. 2019, 9, 8368. [Google Scholar] [CrossRef] [Green Version]
  89. Cunningham, F.; Allen, J.E.; Allen, J.; Alvarez-Jarreta, J.; Amode, M.R.; Armean, I.M.; Austine-Orimoloye, O.; Azov, A.G.; Barnes, I.; Bennett, R.; et al. Ensembl 2022. Nucleic Acids Res. 2022, 50, D988–D995. [Google Scholar] [CrossRef]
  90. Radde, R.; Bolmont, T.; Kaeser, S.A.; Coomaraswamy, J.; Lindau, D.; Stoltze, L.; Calhoun, M.E.; Jäggi, F.; Wolburg, H.; Gengler, S.; et al. Abeta42-driven cerebral amyloidosis in transgenic mice reveals early and robust pathology. EMBO Rep. 2006, 7, 940–946. [Google Scholar] [CrossRef] [Green Version]
  91. Oakley, H.; Cole, S.L.; Logan, S.; Maus, E.; Shao, P.; Craft, J.; Guillozet-Bongaarts, A.; Ohno, M.; Disterhoft, J.; Van Eldik, L.; et al. Intraneuronal beta-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer’s disease mutations: Potential factors in amyloid plaque formation. J. Neurosci. 2006, 26, 10129–10140. [Google Scholar] [CrossRef] [Green Version]
  92. Armbrust, F.; Bickenbach, K.; Marengo, L.; Pietrzik, C.; Becker-Pauly, C. The Swedish dilemma—The almost exclusive use of APPswe-based mouse models impedes adequate evaluation of alternative β-secretases. Biochim. Biophys. Acta Mol. Cell Res. 2022, 1869, 119164. [Google Scholar] [CrossRef] [PubMed]
  93. Kakuda, N.; Funamoto, S.; Yagishita, S.; Takami, M.; Osawa, S.; Dohmae, N.; Ihara, Y. Equimolar production of amyloid β-protein and amyloid precursor protein intracellular domain from β-carboxyl-terminal fragment by γ-secretase. J. Biol. Chem. 2006, 281, 14776–14786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kuhn, A.J.; Raskatov, J. Is the p3 Peptide (Abeta17-40, Abeta17-42) Relevant to the Pathology of Alzheimer’s Disease? J. Alzheimers. Dis. 2020, 74, 43–53. [Google Scholar] [CrossRef] [PubMed]
  95. Bouter, Y.; Dietrich, K.; Wittnam, J.L.; Rezaei-Ghaleh, N.; Pillot, T.; Papot-Couturier, S.; Lefebvre, T.; Sprenger, F.; Wirths, O.; Zweckstetter, M.; et al. N-truncated amyloid beta (Abeta) 4-42 forms stable aggregates and induces acute and long-lasting behavioral deficits. Acta Neuropathol. 2013, 126, 189–205. [Google Scholar] [CrossRef] [Green Version]
  96. LeBlanc, A.C.; Xue, R.; Gambetti, P. Amyloid precursor protein metabolism in primary cell cultures of neurons, astrocytes, and microglia. J. Neurochem. 1996, 66, 2300–2310. [Google Scholar] [CrossRef]
  97. Crescenzi, O.; Tomaselli, S.; Guerrini, R.; Salvadori, S.; D’Ursi, A.M.; Temussi, P.A.; Picone, D. Solution structure of the Alzheimer amyloid β-peptide (1–42) in an apolar microenvironment. Eur. J. Biochem. 2002, 269, 5642–5648. [Google Scholar] [CrossRef]
  98. Oberstein, T.J.; Spitzer, P.; Klafki, H.W.; Linning, P.; Neff, F.; Knolker, H.J.; Lewczuk, P.; Wiltfang, J.; Kornhuber, J.; Maler, J.M. Astrocytes and microglia but not neurons preferentially generate N-terminally truncated Abeta peptides. Neurobiol. Dis. 2015, 73, 24–35. [Google Scholar] [CrossRef] [Green Version]
  99. Bibl, M.; Gallus, M.; Welge, V.; Lehmann, S.; Sparbier, K.; Esselmann, H.; Wiltfang, J. Characterization of cerebrospinal fluid aminoterminally truncated and oxidized amyloid-β peptides. Proteom. Clin. Appl. 2012, 6, 163–169. [Google Scholar] [CrossRef]
  100. Portelius, E.; Price, E.; Brinkmalm, G.; Stiteler, M.; Olsson, M.; Persson, R.; Westman-Brinkmalm, A.; Zetterberg, H.; Simon, A.J.; Blennow, K. A novel pathway for amyloid precursor protein processing. Neurobiol. Aging 2011, 32, 1090–1098. [Google Scholar] [CrossRef]
  101. Marengo, L.; Armbrust, F.; Schoenherr, C.; Storck, S.E.; Schmitt, U.; Zampar, S.; Wirths, O.; Altmeppen, H.; Glatzel, M.; Kaether, C.; et al. Meprin beta knockout reduces brain Abeta levels and rescues learning and memory impairments in the APP/lon mouse model for Alzheimer’s disease. Cell. Mol. Life. Sci. 2022, 79, 168. [Google Scholar] [CrossRef]
  102. Schönherr, C.; Bien, J.; Isbert, S.; Wichert, R.; Prox, J.; Altmeppen, H.; Kumar, S.; Walter, J.; Lichtenthaler, S.F.; Weggen, S.; et al. Generation of aggregation prone N-terminally truncated amyloid β peptides by meprin β depends on the sequence specificity at the cleavage site. Mol. Neurodegener. 2016, 11, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Savastano, A.; Klafki, H.; Haußmann, U.; Oberstein, T.J.; Muller, P.; Wirths, O.; Wiltfang, J.; Bayer, T.A. N-truncated Aβ2-X starting with position two in sporadic Alzheimer’s disease cases and two Alzheimer mouse models. J. Alzheimers Dis. 2016, 49, 101–110. [Google Scholar] [CrossRef] [PubMed]
  104. Wirths, O.; Walter, S.; Kraus, I.; Klafki, H.W.; Stazi, M.; Oberstein, T.J.; Ghiso, J.; Wiltfang, J.; Bayer, T.A.; Weggen, S. N-truncated Aβ4–x peptides in sporadic Alzheimer’s disease cases and transgenic Alzheimer mouse models. Alzheimer’s Res. Ther. 2017, 9, 80. [Google Scholar] [CrossRef] [Green Version]
  105. Walter, S.; Jumpertz, T.; Hüttenrauch, M.; Ogorek, I.; Gerber, H.; Storck, S.E.; Zampar, S.; Dimitrov, M.; Lehmann, S.; Lepka, K.; et al. The metalloprotease ADAMTS4 generates N-truncated Aβ4-x species and marks oligodendrocytes as a source of amyloidogenic peptides in Alzheimer’s disease. Acta Neuropathol. 2019, 137, 239–257. [Google Scholar] [CrossRef] [Green Version]
  106. Marr, R.A.; Guan, H.; Rockenstein, E.; Kindy, M.; Gage, F.H.; Verma, I.; Masliah, E.; Hersh, L.B. Neprilysin regulates amyloid Beta peptide levels. J. Mol. Neurosci. 2004, 22, 5–11. [Google Scholar] [CrossRef]
  107. Pike, C.J.; Overman, M.J.; Cotman, C.W. Amino-terminal deletions enhance aggregation of beta-amyloid peptides in vitro. J. Biol. Chem. 1995, 270, 23895–23898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Rostagno, A.; Cabrera, E.; Lashley, T.; Ghiso, J. N-terminally truncated Aβ4-x proteoforms and their relevance for Alzheimer’s pathophysiology. Transl. Neurodegener. 2022, 11, 30. [Google Scholar] [CrossRef]
  109. Huttenrauch, M.; Brauss, A.; Kurdakova, A.; Borgers, H.; Klinker, F.; Liebetanz, D.; Salinas-Riester, G.; Wiltfang, J.; Klafki, H.W.; Wirths, O. Physical activity delays hippocampal neurodegeneration and rescues memory deficits in an Alzheimer disease mouse model. Transl. Psychiatry 2016, 6, e800. [Google Scholar] [CrossRef] [Green Version]
  110. Sullivan, C.P.; Berg, E.A.; Elliott-Bryant, R.; Fishman, J.B.; McKee, A.C.; Morin, P.J.; Shia, M.A.; Fine, R.E. Pyroglutamate-Aβ 3 and 11 colocalize in amyloid plaques in Alzheimer’s disease cerebral cortex with pyroglutamate-Aβ 11 forming the central core. Neurosci. Lett. 2011, 505, 109–112. [Google Scholar] [CrossRef] [Green Version]
  111. Hao, X.; Zheng, J.; Sun, Y.; Dong, X. Seeding and Cross-Seeding Aggregations of Aβ40 and Its N-Terminal-Truncated Peptide Aβ11–40. Langmuir 2019, 35, 2821–2831. [Google Scholar] [CrossRef]
  112. Barritt, J.D.; Viles, J.H. Truncated Amyloid-beta(11-40/42) from Alzheimer Disease Binds Cu2+ with a Femtomolar Affinity and Influences Fiber Assembly. J. Biol. Chem. 2015, 290, 27791–27802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Busch, L.; al Taleb, Z.; Tsai, Y.L.; Nguyen, V.T.; Lu, Q.; Synatschke, C.V.; Endres, K.; Bufe, B. Amyloid beta and its naturally occurring N-terminal variants are potent activators of human and mouse formyl peptide receptor. J. Biol. Chem. 2022, 102642. [Google Scholar] [CrossRef] [PubMed]
  114. Lalowski, M.; Golabek, A.; Lemere, C.A.; Selkoe, D.J.; Wisniewski, H.M.; Beavis, R.C.; Frangione, B.; Wisniewski, T. The “nonamyloidogenic” p3 fragment (amyloid beta17-42) is a major constituent of Down’s syndrome cerebellar preamyloid. J. Biol. Chem. 1996, 271, 33623–33631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Akiyama, H.; Mori, H.; Saido, T.; Kondo, H.; Ikeda, K.; McGeer, P.L. Occurrence of the diffuse amyloid beta-protein (Abeta) deposits with numerous Abeta-containing glial cells in the cerebral cortex of patients with Alzheimer’s disease. Glia 1999, 25, 324–331. [Google Scholar] [CrossRef]
  116. Moghekar, A.; Rao, S.; Li, M.; Ruben, D.; Mammen, A.; Tang, X.; O’Brien, R.J. Large quantities of Abeta peptide are constitutively released during amyloid precursor protein metabolism in vivo and in vitro. J. Biol. Chem. 2011, 286, 15989–15997. [Google Scholar] [CrossRef] [Green Version]
  117. Hunter, S.; Brayne, C. Do anti-amyloid beta protein antibody cross reactivities confound Alzheimer disease research? J. Negat. Results Biomed. 2017, 16, 1. [Google Scholar] [CrossRef] [Green Version]
  118. Kuhn, A.J.; Raskatov, J.A. A robust preparation method for the amyloidogenic and intrinsically disordered amyloid-alpha peptide. J. Pept. Sci. 2022, e3414. [Google Scholar] [CrossRef]
  119. Dulin, F.; Leveille, F.; Ortega, J.B.; Mornon, J.P.; Buisson, A.; Callebaut, I.; Colloc’h, N. P3 peptide, a truncated form of A beta devoid of synaptotoxic effect, does not assemble into soluble oligomers. FEBS. Lett. 2008, 582, 1865–1870. [Google Scholar] [CrossRef]
  120. Kuhn, A.J.; Abrams, B.S.; Knowlton, S.; Raskatov, J.A. Alzheimer’s Disease “Non-amyloidogenic” p3 Peptide Revisited: A Case for Amyloid-alpha. ACS. Chem. Neurosci. 2020, 11, 1539–1544. [Google Scholar] [CrossRef]
  121. Miller, Y.; Ma, B.; Nussinov, R. Polymorphism of Alzheimer’s Abeta17-42 (p3) oligomers: The importance of the turn location and its conformation. Biophys. J. 2009, 97, 1168–1177. [Google Scholar] [CrossRef]
  122. Streltsov, V.A.; Varghese, J.N.; Masters, C.L.; Nuttall, S.D. Crystal structure of the amyloid-β p3 fragment provides a model for oligomer formation in Alzheimer’s disease. J. Neurosci. 2011, 31, 1419–1426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wei, W.; Norton, D.D.; Wang, X.; Kusiak, J.W. Abeta 17-42 in Alzheimer’s disease activates JNK and caspase-8 leading to neuronal apoptosis. Brain 2002, 125, 2036–2043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Szczepanik, A.M.; Rampe, D.; Ringheim, G.E. Amyloid-beta peptide fragments p3 and p4 induce pro-inflammatory cytokine and chemokine production in vitro and in vivo. J. Neurochem. 2001, 77, 304–317. [Google Scholar] [CrossRef] [PubMed]
  125. Iizuka, T.; Shoji, M.; Harigaya, Y.; Kawarabayashi, T.; Watanabe, M.; Kanai, M.; Hirai, S. Amyloid β-protein ending at Thr43 is a minor component of some diffuse plaques in the Alzheimer’s disease brain, but is not found in cerebrovascular amyloid. Brain. Res. 1995, 702, 275–278. [Google Scholar] [CrossRef]
  126. Jäkel, L.; Boche, D.; Nicoll, J.A.R.; Verbeek, M.M. Aβ43 in human Alzheimer’s disease: Effects of active Aβ42 immunization. Acta Neuropathol. Commun. 2019, 7, 141. [Google Scholar] [CrossRef] [Green Version]
  127. Welander, H.; Frånberg, J.; Graff, C.; Sundström, E.; Winblad, B.; Tjernberg, L.O. Aβ43 is more frequent than Aβ40 in amyloid plaque cores from Alzheimer disease brains. J. Neurochem. 2009, 110, 697–706. [Google Scholar] [CrossRef]
  128. Sandebring, A.; Welander, H.; Winblad, B.; Graff, C.; Tjernberg, L.O. The Pathogenic Aβ43 Is Enriched in Familial and Sporadic Alzheimer Disease. PLoS ONE 2013, 8, e55847. [Google Scholar] [CrossRef]
  129. Kakuda, N.; Takami, M.; Okochi, M.; Kasuga, K.; Ihara, Y.; Ikeuchi, T. Switched Aβ43 generation in familial Alzheimer’s disease with presenilin 1 mutation. Transl. Psychiatry 2021, 11, 558. [Google Scholar] [CrossRef]
  130. Trambauer, J.; Rodríguez Sarmiento, R.M.; Fukumori, A.; Feederle, R.; Baumann, K.; Steiner, H. Aβ43-producing PS1 FAD mutants cause altered substrate interactions and respond to γ-secretase modulation. EMBO Rep. 2020, 21, e47996. [Google Scholar] [CrossRef]
  131. Lauridsen, C.; Sando, S.B.; Møller, I.; Berge, G.; Pomary, P.K.; Grøntvedt, G.R.; Salvesen, Ø.; Bråthen, G.; White, L.R. Cerebrospinal Fluid Aβ43 Is Reduced in Early-Onset Compared to Late-Onset Alzheimer’s Disease, But Has Similar Diagnostic Accuracy to Aβ42. Front. Aging Neurosci. 2017, 9, 210. [Google Scholar] [CrossRef]
  132. Fu, L.; Sun, Y.; Guo, Y.; Chen, Y.; Yu, B.; Zhang, H.; Wu, J.; Yu, X.; Kong, W.; Wu, H. Comparison of neurotoxicity of different aggregated forms of Aβ40, Aβ42 and Aβ43 in cell cultures. J. Pept. Sci. 2017, 23, 245–251. [Google Scholar] [CrossRef] [PubMed]
  133. Saito, T.; Suemoto, T.; Brouwers, N.; Sleegers, K.; Funamoto, S.; Mihira, N.; Matsuba, Y.; Yamada, K.; Nilsson, P.; Takano, J.; et al. Potent amyloidogenicity and pathogenicity of Aβ43. Nat. Neurosci. 2011, 14, 1023–1032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ruiz-Riquelme, A.; Mao, A.; Barghash, M.M.; Lau, H.H.C.; Stuart, E.; Kovacs, G.G.; Nilsson, K.P.R.; Fraser, P.E.; Schmitt-Ulms, G.; Watts, J.C. Aβ43 aggregates exhibit enhanced prion-like seeding activity in mice. Acta Neuropathol. Commun. 2021, 9, 83. [Google Scholar] [CrossRef]
  135. Moore, B.D.; Martin, J.; de Mena, L.; Sanchez, J.; Cruz, P.E.; Ceballos-Diaz, C.; Ladd, T.B.; Ran, Y.; Levites, Y.; Kukar, T.L.; et al. Short Aβ peptides attenuate Aβ42 toxicity in vivo. J. Exp. Med. 2018, 215, 283–301. [Google Scholar] [CrossRef] [PubMed]
  136. Quartey, M.O.; Nyarko, J.N.K.; Maley, J.M.; Barnes, J.R.; Bolanos, M.A.C.; Heistad, R.M.; Knudsen, K.J.; Pennington, P.R.; Buttigieg, J.; De Carvalho, C.E.; et al. The Aβ(1-38) peptide is a negative regulator of the Aβ(1-42) peptide implicated in Alzheimer disease progression. Sci. Rep. 2021, 11, 431. [Google Scholar] [CrossRef]
  137. Wiltfang, J.; Esselmann, H.; Bibl, M.; Smirnov, A.; Otto, M.; Paul, S.; Schmidt, B.; Klafki, H.W.; Maler, M.; Dyrks, T.; et al. Highly conserved and disease-specific patterns of carboxyterminally truncated Abeta peptides 1-37/38/39 in addition to 1-40/42 in Alzheimer’s disease and in patients with chronic neuroinflammation. J. Neurochem. 2002, 81, 481–496. [Google Scholar] [CrossRef]
  138. Cabrera, E.; Mathews, P.; Mezhericher, E.; Beach, T.G.; Deng, J.; Neubert, T.A.; Rostagno, A.; Ghiso, J. Abeta truncated species: Implications for brain clearance mechanisms and amyloid plaque deposition. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 208–225. [Google Scholar] [CrossRef]
  139. Liebsch, F.; Kulic, L.; Teunissen, C.; Shobo, A.; Ulku, I.; Engelschalt, V.; Hancock, M.A.; van der Flier, W.M.; Kunach, P.; Rosa-Neto, P.; et al. Aβ34 is a BACE1-derived degradation intermediate associated with amyloid clearance and Alzheimer’s disease progression. Nat. Commun. 2019, 10, 2240. [Google Scholar] [CrossRef] [Green Version]
  140. Hernandez-Guillamon, M.; Mawhirt, S.; Blais, S.; Montaner, J.; Neubert, T.A.; Rostagno, A.; Ghiso, J. Sequential Amyloid-beta Degradation by the Matrix Metalloproteases MMP-2 and MMP-9. J. Biol. Chem. 2015, 290, 15078–15091. [Google Scholar] [CrossRef] [Green Version]
  141. Caillava, C.; Ranaldi, S.; Lauritzen, I.; Bauer, C.; Fareh, J.; Abraham, J.D.; Checler, F. Study on Aβ34 biology and detection in transgenic mice brains. Neurobiol. Aging 2014, 35, 1570–1581. [Google Scholar] [CrossRef]
  142. Giulian, D.; Haverkamp, L.J.; Yu, J.H.; Karshin, W.; Tom, D.; Li, J.; Kirkpatrick, J.; Kuo, L.M.; Roher, A.E. Specific domains of beta-amyloid from Alzheimer plaque elicit neuron killing in human microglia. J. Neurosci. 1996, 16, 6021–6037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Antonyan, A.; Schlenzig, D.; Schilling, S.; Naumann, M.; Sharoyan, S.; Mardanyan, S.; Demuth, H.U. Concerted action of dipeptidyl peptidase IV and glutaminyl cyclase results in formation of pyroglutamate-modified amyloid peptides in vitro. Neurochem. Int. 2018, 113, 112–119. [Google Scholar] [CrossRef] [PubMed]
  144. Cynis, H.; Schilling, S.; Bodnar, M.; Hoffmann, T.; Heiser, U.; Saido, T.C.; Demuth, H.U. Inhibition of glutaminyl cyclase alters pyroglutamate formation in mammalian cells. Biochim. Biophys. Acta 2006, 1764, 1618–1625. [Google Scholar] [CrossRef] [PubMed]
  145. Jawhar, S.; Wirths, O.; Schilling, S.; Graubner, S.; Demuth, H.U.; Bayer, T.A. Overexpression of glutaminyl cyclase, the enzyme responsible for pyroglutamate A(1) formation, induces behavioral deficits, and glutaminyl cyclase knock-out rescues the behavioral phenotype in 5XFAD mice. J. Biol. Chem. 2011, 286, 4454–4460. [Google Scholar] [CrossRef] [Green Version]
  146. Schilling, S.; Appl, T.; Hoffmann, T.; Cynis, H.; Schulz, K.; Jagla, W.; Friedrich, D.; Wermann, M.; Buchholz, M.; Heiser, U.; et al. Inhibition of glutaminyl cyclase prevents pGlu-Abeta formation after intracortical/hippocampal microinjection in vivo/in situ. J. Neurochem. 2008, 106, 1225–1236. [Google Scholar] [CrossRef]
  147. Schlenzig, D.; Manhart, S.; Cinar, Y.; Kleinschmidt, M.; Hause, G.; Willbold, D.; Funke, S.A.; Schilling, S.; Demuth, H.U. Pyroglutamate formation influences solubility and amyloidogenicity of amyloid peptides. Biochemistry 2009, 48, 7072–7078. [Google Scholar] [CrossRef]
  148. Saido, T.C.; Iwatsubo, T.; Mann, D.M.; Shimada, H.; Ihara, Y.; Kawashima, S. Dominant and differential deposition of distinct beta-amyloid peptide species, A beta N3(pE), in senile plaques. Neuron 1995, 14, 457–466. [Google Scholar] [CrossRef] [Green Version]
  149. Rijal Upadhaya, A.; Kosterin, I.; Kumar, S.; von Arnim, C.A.; Yamaguchi, H.; Fändrich, M.; Walter, J.; Thal, D.R. Biochemical stages of amyloid-β peptide aggregation and accumulation in the human brain and their association with symptomatic and pathologically preclinical Alzheimer’s disease. Brain 2014, 137, 887–903. [Google Scholar] [CrossRef]
  150. Wang, P.N.; Lin, K.J.; Liu, H.C.; Andreasson, U.; Blennow, K.; Zetterberg, H.; Yang, S.Y. Plasma pyroglutamate-modified amyloid beta differentiates amyloid pathology. Alzheimers Dement. 2020, 12, e12029. [Google Scholar] [CrossRef]
  151. Dammers, C.; Schwarten, M.; Buell, A.K.; Willbold, D. Pyroglutamate-modified Abeta(3-42) affects aggregation kinetics of Abeta(1-42) by accelerating primary and secondary pathways. Chem. Sci. 2017, 8, 4996–5004. [Google Scholar] [CrossRef]
  152. Hu, Z.W.; Au, D.F.; Cruceta, L.; Vugmeyster, L.; Qiang, W. N-Terminal Modified Abeta Variants Enable Modulations to the Structures and Cytotoxicity Levels of Wild-Type Abeta Fibrils through Cross-Seeding. ACS Chem. Neurosci. 2020, 11, 2058–2065. [Google Scholar] [CrossRef] [PubMed]
  153. Neddens, J.; Daurer, M.; Flunkert, S.; Beutl, K.; Loeffler, T.; Walker, L.; Attems, J.; Hutter-Paier, B. Correlation of pyroglutamate amyloid beta and ptau Ser202/Thr205 levels in Alzheimer’s disease and related murine models. PLoS ONE 2020, 15, e0235543. [Google Scholar] [CrossRef] [PubMed]
  154. De Kimpe, L.; van Haastert, E.S.; Kaminari, A.; Zwart, R.; Rutjes, H.; Hoozemans, J.J.; Scheper, W. Intracellular accumulation of aggregated pyroglutamate amyloid beta: Convergence of aging and Abeta pathology at the lysosome. Age 2013, 35, 673–687. [Google Scholar] [CrossRef] [Green Version]
  155. Baik, S.H.; Kang, S.; Son, S.M.; Mook-Jung, I. Microglia contributes to plaque growth by cell death due to uptake of amyloid beta in the brain of Alzheimer’s disease mouse model. Glia 2016, 64, 2274–2290. [Google Scholar] [CrossRef] [PubMed]
  156. Joshi, P.; Turola, E.; Ruiz, A.; Bergami, A.; Libera, D.D.; Benussi, L.; Giussani, P.; Magnani, G.; Comi, G.; Legname, G.; et al. Microglia convert aggregated amyloid-beta into neurotoxic forms through the shedding of microvesicles. Cell Death Differ. 2014, 21, 582–593. [Google Scholar] [CrossRef] [Green Version]
  157. Grochowska, K.M.; Yuanxiang, P.; Bar, J.; Raman, R.; Brugal, G.; Sahu, G.; Schweizer, M.; Bikbaev, A.; Schilling, S.; Demuth, H.U.; et al. Posttranslational modification impact on the mechanism by which amyloid-beta induces synaptic dysfunction. EMBO Rep. 2017, 18, 962–981. [Google Scholar] [CrossRef]
  158. Balakrishnan, K.; Rijal Upadhaya, A.; Steinmetz, J.; Reichwald, J.; Abramowski, D.; Fandrich, M.; Kumar, S.; Yamaguchi, H.; Walter, J.; Staufenbiel, M.; et al. Impact of amyloid beta aggregate maturation on antibody treatment in APP23 mice. Acta Neuropathol. Commun. 2015, 3, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Frost, J.L.; Le, K.X.; Cynis, H.; Ekpo, E.; Kleinschmidt, M.; Palmour, R.M.; Ervin, F.R.; Snigdha, S.; Cotman, C.W.; Saido, T.C.; et al. Pyroglutamate-3 amyloid-beta deposition in the brains of humans, non-human primates, canines, and Alzheimer disease-like transgenic mouse models. Am. J. Pathol. 2013, 183, 369–381. [Google Scholar] [CrossRef] [PubMed]
  160. Wirths, O.; Bethge, T.; Marcello, A.; Harmeier, A.; Jawhar, S.; Lucassen, P.J.; Multhaup, G.; Brody, D.L.; Esparza, T.; Ingelsson, M.; et al. Pyroglutamate Abeta pathology in APP/PS1KI mice, sporadic and familial Alzheimer’s disease cases. J. Neural. Transm. 2010, 117, 85–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Crehan, H.; Liu, B.; Kleinschmidt, M.; Rahfeld, J.U.; Le, K.X.; Caldarone, B.J.; Frost, J.L.; Hettmann, T.; Hutter-Paier, B.; O’Nuallain, B.; et al. Effector function of anti-pyroglutamate-3 Aβ antibodies affects cognitive benefit, glial activation and amyloid clearance in Alzheimer’s-like mice. Alzheimers Res. 2020, 12, 12. [Google Scholar] [CrossRef]
  162. Mukherjee, S.; Perez, K.A.; Lago, L.C.; Klatt, S.; McLean, C.A.; Birchall, I.E.; Barnham, K.J.; Masters, C.L.; Roberts, B.R. Quantification of N-terminal amyloid-beta isoforms reveals isomers are the most abundant form of the amyloid-beta peptide in sporadic Alzheimer’s disease. Brain Commun. 2021, 3, fcab028. [Google Scholar] [CrossRef] [PubMed]
  163. Lambeth, T.R.; Riggs, D.L.; Talbert, L.E.; Tang, J.; Coburn, E.; Kang, A.S.; Noll, J.; Augello, C.; Ford, B.D.; Julian, R.R. Spontaneous Isomerization of Long-Lived Proteins Provides a Molecular Mechanism for the Lysosomal Failure Observed in Alzheimer’s Disease. ACS Cent. Sci. 2019, 5, 1387–1395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Fonseca, M.I.; Head, E.; Velazquez, P.; Cotman, C.W.; Tenner, A.J. The presence of isoaspartic acid in beta-amyloid plaques indicates plaque age. Exp. Neurol. 1999, 157, 277–288. [Google Scholar] [CrossRef] [PubMed]
  165. Roher, A.E.; Lowenson, J.D.; Clarke, S.; Wolkow, C.; Wang, R.; Cotter, R.J.; Reardon, I.M.; Zürcher-Neely, H.A.; Heinrikson, R.L.; Ball, M.J. Structural alterations in the peptide backbone of beta-amyloid core protein may account for its deposition and stability in Alzheimer’s disease. J. Biol. Chem. 1993, 268, 3072–3083. [Google Scholar] [CrossRef]
  166. Fabian, H.; Szendrei, G.I.; Mantsch, H.H.; Greenberg, B.D.; ÖTvÖS Jr, L. Synthetic post-translationally modified human Aβ peptide exhibits a markedly increased tendency to form β-pleated sheets in vitro. Eur. J. Biochem. 1994, 221, 959–964. [Google Scholar] [CrossRef]
  167. Szendrei, G.I.; Fabian, H.; Mantsch, H.H.; Lovas, S.; Nyéki, O.; Schön, I.; Otvos, L., Jr. Aspartate-bond isomerization affects the major conformations of synthetic peptides. Eur. J. Biochem. 1994, 226, 917–924. [Google Scholar] [CrossRef] [Green Version]
  168. Barykin, E.P.; Garifulina, A.I.; Kruykova, E.V.; Spirova, E.N.; Anashkina, A.A.; Adzhubei, A.A.; Shelukhina, I.V.; Kasheverov, I.E.; Mitkevich, V.A.; Kozin, S.A.; et al. Isomerization of Asp7 in Beta-Amyloid Enhances Inhibition of the α7 Nicotinic Receptor and Promotes Neurotoxicity. Cells 2019, 8, 771. [Google Scholar] [CrossRef] [Green Version]
  169. Barykin, E.P.; Petrushanko, I.Y.; Burnysheva, K.M.; Makarov, A.A.; Mitkevich, V.A. [Isomerization of Asp7 increases the toxic effects of amyloid beta and its phosphorylated form in SH-SY5Y neuroblastoma cells]. Mol. Biol. 2016, 50, 863–869. [Google Scholar] [CrossRef]
  170. Kuo, Y.M.; Webster, S.; Emmerling, M.R.; De Lima, N.; Roher, A.E. Irreversible dimerization/tetramerization and post-translational modifications inhibit proteolytic degradation of A beta peptides of Alzheimer’s disease. Biochim. Biophys. Acta. 1998, 1406, 291–298. [Google Scholar] [CrossRef] [Green Version]
  171. Kumar, S.; Kapadia, A.; Theil, S.; Joshi, P.; Riffel, F.; Heneka, M.T.; Walter, J. Novel Phosphorylation-State Specific Antibodies Reveal Differential Deposition of Ser26 Phosphorylated Abeta Species in a Mouse Model of Alzheimer’s Disease. Front. Mol. Neurosci. 2020, 13, 619639. [Google Scholar] [CrossRef]
  172. Kumar, S.; Wirths, O.; Stüber, K.; Wunderlich, P.; Koch, P.; Theil, S.; Rezaei-Ghaleh, N.; Zweckstetter, M.; Bayer, T.A.; Brüstle, O.; et al. Phosphorylation of the amyloid β-peptide at Ser26 stabilizes oligomeric assembly and increases neurotoxicity. Acta Neuropathol. 2016, 131, 525–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Ashby, E.L.; Miners, J.S.; Kumar, S.; Walter, J.; Love, S.; Kehoe, P.G. Investigation of Abeta phosphorylated at serine 8 (pAbeta) in Alzheimer’s disease, dementia with Lewy bodies and vascular dementia. Neuropathol. Appl. Neurobiol. 2015, 41, 428–444. [Google Scholar] [CrossRef] [PubMed]
  174. Kumar, S.; Singh, S.; Hinze, D.; Josten, M.; Sahl, H.G.; Siepmann, M.; Walter, J. Phosphorylation of amyloid-β peptide at serine 8 attenuates its clearance via insulin-degrading and angiotensin-converting enzymes. J. Biol. Chem. 2012, 287, 8641–8651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Rezaei-Ghaleh, N.; Amininasab, M.; Kumar, S.; Walter, J.; Zweckstetter, M. Phosphorylation modifies the molecular stability of beta-amyloid deposits. Nat. Commun. 2016, 7, 11359. [Google Scholar] [CrossRef] [Green Version]
  176. Friedemann, M.; Helk, E.; Tiiman, A.; Zovo, K.; Palumaa, P.; Tougu, V. Effect of methionine-35 oxidation on the aggregation of amyloid-beta peptide. Biochem. Biophys. Rep. 2015, 3, 94–99. [Google Scholar] [CrossRef] [Green Version]
  177. Razzokov, J.; Yusupov, M.; Bogaerts, A. Oxidation destabilizes toxic amyloid beta peptide aggregation. Sci. Rep. 2019, 9, 5476. [Google Scholar] [CrossRef] [Green Version]
  178. Head, E.; Garzon-Rodriguez, W.; Johnson, J.K.; Lott, I.T.; Cotman, C.W.; Glabe, C. Oxidation of Aβ and Plaque Biogenesis in Alzheimer’s Disease and Down Syndrome. Neurobiol. Dis. 2001, 8, 792–806. [Google Scholar] [CrossRef] [Green Version]
  179. Tomiyama, T.; Asano, S.; Furiya, Y.; Shirasawa, T.; Endo, N.; Mori, H. Racemization of Asp23 residue affects the aggregation properties of Alzheimer amyloid beta protein analogues. J. Biol. Chem. 1994, 269, 10205–10208. [Google Scholar] [CrossRef]
  180. Kang, J.; Lemaire, H.G.; Unterbeck, A.; Salbaum, J.M.; Masters, C.L.; Grzeschik, K.H.; Multhaup, G.; Beyreuther, K.; Müller-Hill, B. The precursor of Alzheimer’s disease amyloid A4 protein resembles a cell-surface receptor. Nature 1987, 325, 733–736. [Google Scholar] [CrossRef]
  181. Nalivaeva, N.N.; Turner, A.J. The amyloid precursor protein: A biochemical enigma in brain development, function and disease. FEBS Lett. 2013, 587, 2046–2054. [Google Scholar] [CrossRef]
  182. Panegyres, P.K.; Zafiris-Toufexis, K.; Kakulas, B.A. Amyloid precursor protein gene isoforms in Alzheimer’s disease and other neurodegenerative disorders. J. Neurol. Sci. 2000, 173, 81–92. [Google Scholar] [CrossRef]
  183. Beckmann, A.M.; Glebov, K.; Walter, J.; Merkel, O.; Mangold, M.; Schmidt, F.; Becker-Pauly, C.; Gütschow, M.; Stirnberg, M. The intact Kunitz domain protects the amyloid precursor protein from being processed by matriptase-2. Biol. Chem. 2016, 397, 777–790. [Google Scholar] [CrossRef] [PubMed]
  184. Dawkins, E.; Small, D.H. Insights into the physiological function of the β-amyloid precursor protein: Beyond Alzheimer’s disease. J. Neurochem. 2014, 129, 756–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Zhao, J.; Paganini, L.; Mucke, L.; Gordon, M.; Refolo, L.; Carman, M.; Sinha, S.; Oltersdorf, T.; Lieberburg, I.; McConlogue, L. Beta-secretase processing of the beta-amyloid precursor protein in transgenic mice is efficient in neurons but inefficient in astrocytes. J. Biol. Chem. 1996, 271, 31407–31411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Cai, H.; Wang, Y.; McCarthy, D.; Wen, H.; Borchelt, D.R.; Price, D.L.; Wong, P.C. BACE1 is the major beta-secretase for generation of Abeta peptides by neurons. Nat. Neurosci. 2001, 4, 233–234. [Google Scholar] [CrossRef]
  187. Koo, E.H.; Sisodia, S.S.; Archer, D.R.; Martin, L.J.; Weidemann, A.; Beyreuther, K.; Fischer, P.; Masters, C.L.; Price, D.L. Precursor of amyloid protein in Alzheimer disease undergoes fast anterograde axonal transport. Proc. Natl. Acad. Sci. USA 1990, 87, 1561–1565. [Google Scholar] [CrossRef] [Green Version]
  188. Kang, J.; Muller-Hill, B. Differential splicing of Alzheimer’s disease amyloid A4 precursor RNA in rat tissues: PreA4(695) mRNA is predominantly produced in rat and human brain. Biochem. Biophys. Res. Commun. 1990, 166, 1192–1200. [Google Scholar] [CrossRef]
  189. LeBlanc, A.C.; Chen, H.Y.; Autilio-Gambetti, L.; Gambetti, P. Differential APP gene expression in rat cerebral cortex, meninges, and primary astroglial, microglial and neuronal cultures. FEBS Lett. 1991, 292, 171–178. [Google Scholar] [CrossRef] [Green Version]
  190. LeBlanc, A.C.; Papadopoulos, M.; Belair, C.; Chu, W.; Crosato, M.; Powell, J.; Goodyer, C.G. Processing of amyloid precursor protein in human primary neuron and astrocyte cultures. J. Neurochem. 1997, 68, 1183–1190. [Google Scholar] [CrossRef]
  191. Laird, F.M.; Cai, H.; Savonenko, A.V.; Farah, M.H.; He, K.; Melnikova, T.; Wen, H.; Chiang, H.C.; Xu, G.; Koliatsos, V.E.; et al. BACE1, a major determinant of selective vulnerability of the brain to amyloid-beta amyloidogenesis, is essential for cognitive, emotional, and synaptic functions. J. Neurosci. 2005, 25, 11693–11709. [Google Scholar] [CrossRef]
  192. Haass, C.; Hung, A.Y.; Selkoe, D.J. Processing of beta-amyloid precursor protein in microglia and astrocytes favors an internal localization over constitutive secretion. J. Neurosci. 1991, 11, 3783–3793. [Google Scholar] [CrossRef] [PubMed]
  193. Glenner, G.G.; Wong, C.W. Alzheimer’s disease: Initial report of the purification and characterization of a novel cerebrovascular amyloid protein. Biochem. Biophys. Res. Commun. 1984, 120, 885–890. [Google Scholar] [CrossRef]
  194. Sergeant, N.; Bombois, S.; Ghestem, A.; Drobecq, H.; Kostanjevecki, V.; Missiaen, C.; Wattez, A.; David, J.P.; Vanmechelen, E.; Sergheraert, C.; et al. Truncated beta-amyloid peptide species in pre-clinical Alzheimer’s disease as new targets for the vaccination approach. J. Neurochem. 2003, 85, 1581–1591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Oberstein, T.J.; Utz, J.; Spitzer, P.; Klafki, H.W.; Wiltfang, J.; Lewczuk, P.; Kornhuber, J.; Maler, J.M. The Role of Cathepsin B in the Degradation of Abeta and in the Production of Abeta Peptides Starting With Ala2 in Cultured Astrocytes. Front. Mol. Neurosci. 2020, 13, 615740. [Google Scholar] [CrossRef] [PubMed]
  196. Bien, J.; Jefferson, T.; Causevic, M.; Jumpertz, T.; Munter, L.; Multhaup, G.; Weggen, S.; Becker-Pauly, C.; Pietrzik, C.U. The metalloprotease meprin beta generates amino terminal-truncated amyloid beta peptide species. J. Biol. Chem. 2012, 287, 33304–33313. [Google Scholar] [CrossRef] [Green Version]
  197. Howell, S.; Nalbantoglu, J.; Crine, P. Neutral endopeptidase can hydrolyze beta-amyloid(1-40) but shows no effect on beta-amyloid precursor protein metabolism. Peptides 1995, 16, 647–652. [Google Scholar] [CrossRef]
  198. Liao, M.C.; Ahmed, M.; Smith, S.O.; Van Nostrand, W.E. Degradation of amyloid beta protein by purified myelin basic protein. J. Biol. Chem. 2009, 284, 28917–28925. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Sevalle, J.; Amoyel, A.; Robert, P.; Fournie-Zaluski, M.C.; Roques, B.; Checler, F. Aminopeptidase A contributes to the N-terminal truncation of amyloid beta-peptide. J. Neurochem. 2009, 109, 248–256. [Google Scholar] [CrossRef] [PubMed]
  200. Bayer, T.A.; Wirths, O. Focusing the amyloid cascade hypothesis on N-truncated Abeta peptides as drug targets against Alzheimer’s disease. Acta Neuropathol. 2014, 127, 787–801. [Google Scholar] [CrossRef] [Green Version]
  201. Malhotra, S.K.; Shnitka, T.K.; Elbrink, J. Reactive astrocytes--a review. Cytobios. 1990, 61, 133–160. [Google Scholar]
  202. Siman, R.; Card, J.P.; Nelson, R.B.; Davis, L.G. Expression of beta-amyloid precursor protein in reactive astrocytes following neuronal damage. Neuron 1989, 3, 275–285. [Google Scholar] [CrossRef]
  203. Mita, S.; Schon, E.A.; Herbert, J. Widespread expression of amyloid beta-protein precursor gene in rat brain. Am. J. Pathol. 1989, 134, 1253–1261. [Google Scholar] [PubMed]
  204. Zhao, J.; O’Connor, T.; Vassar, R. The contribution of activated astrocytes to Abeta production: Implications for Alzheimer’s disease pathogenesis. J. Neuroinflammation. 2011, 8, 150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Banati, R.B.; Gehrmann, J.; Czech, C.; Monning, U.; Jones, L.L.; Konig, G.; Beyreuther, K.; Kreutzberg, G.W. Early and rapid de novo synthesis of Alzheimer beta A4-amyloid precursor protein (APP) in activated microglia. Glia 1993, 9, 199–210. [Google Scholar] [CrossRef]
  206. Bauer, J.; Konig, G.; Strauss, S.; Jonas, U.; Ganter, U.; Weidemann, A.; Monning, U.; Masters, C.L.; Volk, B.; Berger, M.; et al. In-vitro matured human macrophages express Alzheimer’s beta A4-amyloid precursor protein indicating synthesis in microglial cells. FEBS Lett. 1991, 282, 335–340. [Google Scholar] [CrossRef] [Green Version]
  207. Monning, U.; Sandbrink, R.; Weidemann, A.; Banati, R.B.; Masters, C.L.; Beyreuther, K. Extracellular matrix influences the biogenesis of amyloid precursor protein in microglial cells. J. Biol. Chem. 1995, 270, 7104–7110. [Google Scholar] [CrossRef] [Green Version]
  208. Singh, N.; Das, B.; Zhou, J.; Hu, X.; Yan, R. Targeted BACE-1 inhibition in microglia enhances amyloid clearance and improved cognitive performance. Sci. Adv. 2022, 8, eabo3610. [Google Scholar] [CrossRef]
  209. Gunner, G.; Cheadle, L.; Johnson, K.M.; Ayata, P.; Badimon, A.; Mondo, E.; Nagy, M.A.; Liu, L.; Bemiller, S.M.; Kim, K.W.; et al. Sensory lesioning induces microglial synapse elimination via ADAM10 and fractalkine signaling. Nat. Neurosci. 2019, 22, 1075–1088. [Google Scholar] [CrossRef]
  210. Nadler, Y.; Alexandrovich, A.; Grigoriadis, N.; Hartmann, T.; Rao, K.S.; Shohami, E.; Stein, R. Increased expression of the gamma-secretase components presenilin-1 and nicastrin in activated astrocytes and microglia following traumatic brain injury. Glia 2008, 56, 552–567. [Google Scholar] [CrossRef]
  211. Bitting, L.; Naidu, A.; Cordell, B.; Murphy, G.M., Jr. Beta-amyloid peptide secretion by a microglial cell line is induced by beta-amyloid-(25-35) and lipopolysaccharide. J. Biol. Chem. 1996, 271, 16084–16089. [Google Scholar] [CrossRef] [Green Version]
  212. Austin, S.A.; Sens, M.A.; Combs, C.K. Amyloid Precursor Protein Mediates a Tyrosine Kinase-Dependent Activation Response in Endothelial Cells. J. Neurosci. 2009, 29, 14451–14462. [Google Scholar] [CrossRef] [PubMed]
  213. Sondag, C.M.; Combs, C.K. Amyloid precursor protein mediates proinflammatory activation of monocytic lineage cells. J. Biol. Chem. 2004, 279, 14456–14463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Manocha, G.D.; Floden, A.M.; Rausch, K.; Kulas, J.A.; McGregor, B.A.; Rojanathammanee, L.; Puig, K.R.; Puig, K.L.; Karki, S.; Nichols, M.R.; et al. APP Regulates Microglial Phenotype in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2016, 36, 8471–8486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Garcia-Ladona, F.J.; Huss, Y.; Frey, P.; Ghandour, M.S. Oligodendrocytes express different isoforms of beta-amyloid precursor protein in chemically defined cell culture conditions: In Situ hybridization and immunocytochemical detection. J. Neurosci. Res. 1997, 50, 50–61. [Google Scholar] [CrossRef]
  216. Sandbrink, R.; Masters, C.L.; Beyreuther, K. APP gene family. Alternative splicing generates functionally related isoforms. Ann. N. Y. Acad. Sci. 1996, 777, 281–287. [Google Scholar] [CrossRef]
  217. Mizuguchi, M.; Ikeda, K.; Kim, S.U. Differential distribution of cellular forms of beta-amyloid precursor protein in murine glial cell cultures. Brain. Res. 1992, 584, 219–225. [Google Scholar] [CrossRef]
  218. Nihonmatsu-Kikuchi, N.; Yu, X.J.; Matsuda, Y.; Ozawa, N.; Ito, T.; Satou, K.; Kaname, T.; Iwasaki, Y.; Akagi, A.; Yoshida, M.; et al. Essential roles of plexin-B3(+) oligodendrocyte precursor cells in the pathogenesis of Alzheimer’s disease. Commun. Biol. 2021, 4, 870. [Google Scholar] [CrossRef]
  219. Card, J.P.; Meade, R.P.; Davis, L.G. Immunocytochemical localization of the precursor protein for beta-amyloid in the rat central nervous system. Neuron 1988, 1, 835–846. [Google Scholar] [CrossRef]
  220. Kawarabayashi, T.; Shoji, M.; Harigaya, Y.; Yamaguchi, H.; Hirai, S. Amyloid beta/A4 protein precursor is widely distributed in both the central and peripheral nervous systems of the mouse. Brain Res. 1991, 552, 1–7. [Google Scholar] [CrossRef]
  221. Palacios, G.; Palacios, J.M.; Mengod, G.; Frey, P. Beta-amyloid precursor protein localization in the Golgi apparatus in neurons and oligodendrocytes. An immunocytochemical structural and ultrastructural study in normal and axotomized neurons. Brain. Res. Mol. Brain Res. 1992, 15, 195–206. [Google Scholar] [CrossRef]
  222. Mizuguchi, M.; Ikeda, K.; Kim, S.U. beta-Amyloid precursor protein of Alzheimer’s disease in cultured bovine oligodendrocytes. J. Neurosci. Res. 1992, 32, 34–42. [Google Scholar] [CrossRef] [PubMed]
  223. Lin, J.; Luo, J.; Redies, C. Differential expression of five members of the ADAM family in the developing chicken brain. Neuroscience 2008, 157, 360–375. [Google Scholar] [CrossRef]
  224. Jung, M.; Krämer, E.; Grzenkowski, M.; Tang, K.; Blakemore, W.; Aguzzi, A.; Khazaie, K.; Chlichlia, K.; von Blankenfeld, G.; Kettenmann, H.; et al. Lines of murine oligodendroglial precursor cells immortalized by an activated neu tyrosine kinase show distinct degrees of interaction with axons in vitro and in vivo. Eur. J. Neurosci. 1995, 7, 1245–1265. [Google Scholar] [CrossRef]
  225. Sakry, D.; Neitz, A.; Singh, J.; Frischknecht, R.; Marongiu, D.; Biname, F.; Perera, S.S.; Endres, K.; Lutz, B.; Radyushkin, K.; et al. Oligodendrocyte precursor cells modulate the neuronal network by activity-dependent ectodomain cleavage of glial NG2. PLoS Biol. 2014, 12, e1001993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Ries, M.; Sastre, M. Mechanisms of Aβ Clearance and Degradation by Glial Cells. Front. Aging Neurosci. 2016, 8, 160. [Google Scholar] [CrossRef] [Green Version]
  227. Brandenburg, L.O.; Konrad, M.; Wruck, C.; Koch, T.; Pufe, T.; Lucius, R. Involvement of formyl-peptide-receptor-like-1 and phospholipase D in the internalization and signal transduction of amyloid beta 1-42 in glial cells. Neuroscience 2008, 156, 266–276. [Google Scholar] [CrossRef] [PubMed]
  228. Kong, Y.; Ruan, L.; Qian, L.; Liu, X.; Le, Y. Norepinephrine promotes microglia to uptake and degrade amyloid beta peptide through upregulation of mouse formyl peptide receptor 2 and induction of insulin-degrading enzyme. J. Neurosci. 2010, 30, 11848–11857. [Google Scholar] [CrossRef] [Green Version]
  229. Yazawa, H.; Yu, Z.X.; Takeda, L.Y.; Gong, W.; Ferrans, V.J.; Oppenheim, J.J.; Li, C.C.; Wang, J.M. Beta amyloid peptide (Abeta42) is internalized via the G-protein-coupled receptor FPRL1 and forms fibrillar aggregates in macrophages. FASEB J. 2001, 15, 2454–2462. [Google Scholar] [CrossRef] [PubMed]
  230. Heurtaux, T.; Michelucci, A.; Losciuto, S.; Gallotti, C.; Felten, P.; Dorban, G.; Grandbarbe, L.; Morga, E.; Heuschling, P. Microglial activation depends on beta-amyloid conformation: Role of the formylpeptide receptor 2. J. Neurochem. 2010, 114, 576–586. [Google Scholar] [CrossRef]
  231. Le, Y.; Gong, W.; Tiffany, H.L.; Tumanov, A.; Nedospasov, S.; Shen, W.; Dunlop, N.M.; Gao, J.L.; Murphy, P.M.; Oppenheim, J.J.; et al. Amyloid (beta)42 activates a G-protein-coupled chemoattractant receptor, FPR-like-1. J. Neurosci. 2001, 21, Rc123. [Google Scholar] [CrossRef]
  232. Zhang, H.; Su, Y.J.; Zhou, W.W.; Wang, S.W.; Xu, P.X.; Yu, X.L.; Liu, R.T. Activated scavenger receptor A promotes glial internalization of aβ. PLoS ONE 2014, 9, e94197. [Google Scholar] [CrossRef] [PubMed]
  233. Husemann, J.; Loike, J.D.; Kodama, T.; Silverstein, S.C. Scavenger receptor class B type I (SR-BI) mediates adhesion of neonatal murine microglia to fibrillar beta-amyloid. J. Neuroimmunol. 2001, 114, 142–150. [Google Scholar] [CrossRef]
  234. Coraci, I.S.; Husemann, J.; Berman, J.W.; Hulette, C.; Dufour, J.H.; Campanella, G.K.; Luster, A.D.; Silverstein, S.C.; El-Khoury, J.B. CD36, a class B scavenger receptor, is expressed on microglia in Alzheimer’s disease brains and can mediate production of reactive oxygen species in response to beta-amyloid fibrils. Am. J. Pathol. 2002, 160, 101–112. [Google Scholar] [CrossRef]
  235. El Khoury, J.B.; Moore, K.J.; Means, T.K.; Leung, J.; Terada, K.; Toft, M.; Freeman, M.W.; Luster, A.D. CD36 mediates the innate host response to beta-amyloid. J. Exp. Med. 2003, 197, 1657–1666. [Google Scholar] [CrossRef] [Green Version]
  236. Liu, S.; Liu, Y.; Hao, W.; Wolf, L.; Kiliaan, A.J.; Penke, B.; Rube, C.E.; Walter, J.; Heneka, M.T.; Hartmann, T.; et al. TLR2 is a primary receptor for Alzheimer’s amyloid beta peptide to trigger neuroinflammatory activation. J. Immunol. 2012, 188, 1098–1107. [Google Scholar] [CrossRef] [Green Version]
  237. Richard, K.L.; Filali, M.; Prefontaine, P.; Rivest, S. Toll-like receptor 2 acts as a natural innate immune receptor to clear amyloid beta 1-42 and delay the cognitive decline in a mouse model of Alzheimer’s disease. J. Neurosci. 2008, 28, 5784–5793. [Google Scholar] [CrossRef] [Green Version]
  238. Tahara, K.; Kim, H.D.; Jin, J.J.; Maxwell, J.A.; Li, L.; Fukuchi, K.I. Role of toll-like receptor signalling in A uptake and clearance. Brain 2006, 129, 3006–3019. [Google Scholar] [CrossRef] [Green Version]
  239. Li, H.Q.; Chen, C.; Dou, Y.; Wu, H.J.; Liu, Y.J.; Lou, H.F.; Zhang, J.M.; Li, X.M.; Wang, H.; Duan, S. P2Y4 receptor-mediated pinocytosis contributes to amyloid beta-induced self-uptake by microglia. Mol. Cell. Biol. 2013, 33, 4282–4293. [Google Scholar] [CrossRef] [Green Version]
  240. Nazere, K.; Takahashi, T.; Hara, N.; Muguruma, K.; Nakamori, M.; Yamazaki, Y.; Morino, H.; Maruyama, H. Amyloid Beta Is Internalized via Macropinocytosis, an HSPG- and Lipid Raft-Dependent and Rac1-Mediated Process. Front. Mol. Neurosci. 2022, 15. [Google Scholar] [CrossRef]
  241. Cho, M.H.; Cho, K.; Kang, H.J.; Jeon, E.Y.; Kim, H.S.; Kwon, H.J.; Kim, H.M.; Kim, D.H.; Yoon, S.Y. Autophagy in microglia degrades extracellular beta-amyloid fibrils and regulates the NLRP3 inflammasome. Autophagy 2014, 10, 1761–1775. [Google Scholar] [CrossRef] [Green Version]
  242. Prakash, P.; Jethava, K.P.; Korte, N.; Izquierdo, P.; Favuzzi, E.; Rose, I.V.L.; Guttenplan, K.A.; Manchanda, P.; Dutta, S.; Rochet, J.C.; et al. Monitoring phagocytic uptake of amyloid beta into glial cell lysosomes in real time. Chem. Sci. 2021, 12, 10901–10918. [Google Scholar] [CrossRef] [PubMed]
  243. Yim, W.W.-Y.; Mizushima, N. Lysosome biology in autophagy. Cell. Discov. 2020, 6, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. He, W.T.; Wan, H.; Hu, L.; Chen, P.; Wang, X.; Huang, Z.; Yang, Z.H.; Zhong, C.Q.; Han, J. Gasdermin D is an executor of pyroptosis and required for interleukin-1beta secretion. Cell. Res. 2015, 25, 1285–1298. [Google Scholar] [CrossRef] [PubMed]
  245. Sosna, J.; Philipp, S.; Albay, R., 3rd; Reyes-Ruiz, J.M.; Baglietto-Vargas, D.; LaFerla, F.M.; Glabe, C.G. Early long-term administration of the CSF1R inhibitor PLX3397 ablates microglia and reduces accumulation of intraneuronal amyloid, neuritic plaque deposition and pre-fibrillar oligomers in 5XFAD mouse model of Alzheimer’s disease. Mol. Neurodegener. 2018, 13, 11. [Google Scholar] [CrossRef]
  246. Spangenberg, E.; Severson, P.L.; Hohsfield, L.A.; Crapser, J.; Zhang, J.; Burton, E.A.; Zhang, Y.; Spevak, W.; Lin, J.; Phan, N.Y.; et al. Sustained microglial depletion with CSF1R inhibitor impairs parenchymal plaque development in an Alzheimer’s disease model. Nat. Commun. 2019, 10, 3758. [Google Scholar] [CrossRef] [Green Version]
  247. Pomilio, C.; Gorojod, R.M.; Riudavets, M.; Vinuesa, A.; Presa, J.; Gregosa, A.; Bentivegna, M.; Alaimo, A.; Alcon, S.P.; Sevlever, G.; et al. Microglial autophagy is impaired by prolonged exposure to beta-amyloid peptides: Evidence from experimental models and Alzheimer’s disease patients. Geroscience 2020, 42, 613–632. [Google Scholar] [CrossRef]
  248. Liang, T.; Zhang, Y.; Wu, S.; Chen, Q.; Wang, L. The Role of NLRP3 Inflammasome in Alzheimer’s Disease and Potential Therapeutic Targets. Front. Pharm. 2022, 13, 845185. [Google Scholar] [CrossRef]
  249. Swanson, K.V.; Deng, M.; Ting, J.P.Y. The NLRP3 inflammasome: Molecular activation and regulation to therapeutics. Nat. Rev. Immunol. 2019, 19, 477–489. [Google Scholar] [CrossRef]
  250. Bauernfeind, F.G.; Horvath, G.; Stutz, A.; Alnemri, E.S.; MacDonald, K.; Speert, D.; Fernandes-Alnemri, T.; Wu, J.; Monks, B.G.; Fitzgerald, K.A.; et al. Cutting edge: NF-kappaB activating pattern recognition and cytokine receptors license NLRP3 inflammasome activation by regulating NLRP3 expression. J. Immunol. 2009, 183, 787–791. [Google Scholar] [CrossRef] [Green Version]
  251. Sheedy, F.J.; Grebe, A.; Rayner, K.J.; Kalantari, P.; Ramkhelawon, B.; Carpenter, S.B.; Becker, C.E.; Ediriweera, H.N.; Mullick, A.E.; Golenbock, D.T.; et al. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat. Immunol. 2013, 14, 812–820. [Google Scholar] [CrossRef] [Green Version]
  252. D’Amico, R.; Fusco, R.; Cordaro, M.; Siracusa, R.; Peritore, A.F.; Gugliandolo, E.; Crupi, R.; Scuto, M.; Cuzzocrea, S.; Di Paola, R.; et al. Modulation of NLRP3 Inflammasome through Formyl Peptide Receptor 1 (Fpr-1) Pathway as a New Therapeutic Target in Bronchiolitis Obliterans Syndrome. Int. J. Mol. Sci. 2020, 2144. [Google Scholar] [CrossRef] [PubMed]
  253. Nakanishi, A.; Kaneko, N.; Takeda, H.; Sawasaki, T.; Morikawa, S.; Zhou, W.; Kurata, M.; Yamamoto, T.; Akbar, S.M.F.; Zako, T.; et al. Amyloid beta directly interacts with NLRP3 to initiate inflammasome activation: Identification of an intrinsic NLRP3 ligand in a cell-free system. Inflamm. Regen. 2018, 38, 27. [Google Scholar] [CrossRef] [PubMed]
  254. Venegas, C.; Kumar, S.; Franklin, B.S.; Dierkes, T.; Brinkschulte, R.; Tejera, D.; Vieira-Saecker, A.; Schwartz, S.; Santarelli, F.; Kummer, M.P.; et al. Microglia-derived ASC specks cross-seed amyloid-β in Alzheimer’s disease. Nature 2017, 552, 355–361. [Google Scholar] [CrossRef] [PubMed]
  255. Griffin, W.S.; Sheng, J.G.; Roberts, G.W.; Mrak, R.E. Interleukin-1 expression in different plaque types in Alzheimer’s disease: Significance in plaque evolution. J. Neuropathol. Exp. Neurol. 1995, 54, 276–281. [Google Scholar] [CrossRef]
  256. Ciallella, J.R.; Ikonomovic, M.D.; Paljug, W.R.; Wilbur, Y.I.; Dixon, C.E.; Kochanek, P.M.; Marion, D.W.; DeKosky, S.T. Changes in expression of amyloid precursor protein and interleukin-1beta after experimental traumatic brain injury in rats. J. Neurotrauma. 2002, 19, 1555–1567. [Google Scholar] [CrossRef]
  257. Donnelly, R.J.; Friedhoff, A.J.; Beer, B.; Blume, A.J.; Vitek, M.P. Interleukin-1 stimulates the beta-amyloid precursor protein promoter. Cell. Mol. Neurobiol. 1990, 10, 485–495. [Google Scholar] [CrossRef]
  258. Buxbaum, J.D.; Oishi, M.; Chen, H.I.; Pinkas-Kramarski, R.; Jaffe, E.A.; Gandy, S.E.; Greengard, P. Cholinergic agonists and interleukin 1 regulate processing and secretion of the Alzheimer beta/A4 amyloid protein precursor. Proc. Natl. Acad. Sci. USA 1992, 89, 10075–10078. [Google Scholar] [CrossRef] [Green Version]
  259. Kong, Q.; Peterson, T.S.; Baker, O.; Stanley, E.; Camden, J.; Seye, C.I.; Erb, L.; Simonyi, A.; Wood, W.G.; Sun, G.Y.; et al. Interleukin-1beta enhances nucleotide-induced and alpha-secretase-dependent amyloid precursor protein processing in rat primary cortical neurons via up-regulation of the P2Y(2) receptor. J. Neurochem. 2009, 109, 1300–1310. [Google Scholar] [CrossRef] [Green Version]
  260. Tachida, Y.; Nakagawa, K.; Saito, T.; Saido, T.C.; Honda, T.; Saito, Y.; Murayama, S.; Endo, T.; Sakaguchi, G.; Kato, A.; et al. Interleukin-1β up-regulates TACE to enhance α-cleavage of APP in neurons: Resulting decrease in Aβ production. J. Neurochem. 2008, 104, 1387–1393. [Google Scholar] [CrossRef]
  261. Griffin, W.S.; Stanley, L.C.; Ling, C.; White, L.; MacLeod, V.; Perrot, L.J.; White, C.L., 3rd; Araoz, C. Brain interleukin 1 and S-100 immunoreactivity are elevated in Down syndrome and Alzheimer disease. Proc. Natl. Acad. Sci. USA 1989, 86, 7611–7615. [Google Scholar] [CrossRef] [Green Version]
  262. Heneka, M.T.; Kummer, M.P.; Stutz, A.; Delekate, A.; Schwartz, S.; Vieira-Saecker, A.; Griep, A.; Axt, D.; Remus, A.; Tzeng, T.C.; et al. NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature 2013, 493, 674–678. [Google Scholar] [CrossRef] [PubMed]
  263. Shaftel, S.S.; Kyrkanides, S.; Olschowka, J.A.; Miller, J.N.; Johnson, R.E.; O’Banion, M.K. Sustained hippocampal IL-1 beta overexpression mediates chronic neuroinflammation and ameliorates Alzheimer plaque pathology. J. Clin. Invest. 2007, 117, 1595–1604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Simpson, D.S.A.; Oliver, P.L. ROS Generation in Microglia: Understanding Oxidative Stress and Inflammation in Neurodegenerative Disease. Antioxidants 2020, 9, 743. [Google Scholar] [CrossRef]
  265. Zhu, M.; Li, D.; Wu, Y.; Huang, X.; Wu, M. TREM-2 promotes macrophage-mediated eradication of Pseudomonas aeruginosa via a PI3K/Akt pathway. Scand. J. Immunol. 2014, 79, 187–196. [Google Scholar] [CrossRef] [PubMed]
  266. Tiffany, H.L.; Lavigne, M.C.; Cui, Y.H.; Wang, J.M.; Leto, T.L.; Gao, J.L.; Murphy, P.M. Amyloid-beta induces chemotaxis and oxidant stress by acting at formylpeptide receptor 2, a G protein-coupled receptor expressed in phagocytes and brain. J. Biol. Chem. 2001, 276, 23645–23652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Aoyama, K. Glutathione in the Brain. Int. J. Mol. Sci. 2021, 5010. [Google Scholar] [CrossRef] [PubMed]
  268. Skoumalová, A.; Hort, J. Blood markers of oxidative stress in Alzheimer’s disease. J. Cell. Mol. Med. 2012, 16, 2291–2300. [Google Scholar] [CrossRef]
  269. Galasko, D.; Montine, T.J. Biomarkers of oxidative damage and inflammation in Alzheimer’s disease. Biomark. Med. 2010, 4, 27–36. [Google Scholar] [CrossRef] [Green Version]
  270. Hatanaka, H.; Hanyu, H.; Hirose, D.; Fukusawa, R.; Namioka, N.; Iwamoto, T. Peripheral Oxidative Stress Markers in Individuals with Alzheimer’s Disease with or without Cerebrovascular Disease. J. Am. Geriatr. Soc. 2015, 63, 1472–1474. [Google Scholar] [CrossRef]
  271. Sakakibara, R.; Kawai, T. Cerebrospinal fluid oxidative stress markers in Alzheimer’s disease. Neurol. Clin. Neurosci. 2020, 8, 232–240. [Google Scholar] [CrossRef]
  272. Bruce-Keller, A.J.; Gupta, S.; Parrino, T.E.; Knight, A.G.; Ebenezer, P.J.; Weidner, A.M.; LeVine, H., 3rd; Keller, J.N.; Markesbery, W.R. NOX activity is increased in mild cognitive impairment. Antioxid. Redox. Signal. 2010, 12, 1371–1382. [Google Scholar] [CrossRef] [PubMed]
  273. Geng, L.; Fan, L.M.; Liu, F.; Smith, C.; Li, J.M. Nox2 dependent redox-regulation of microglial response to amyloid-β stimulation and microgliosis in aging. Sci. Rep. 2020, 10, 1582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Bourdel-Marchasson, I.; Delmas-Beauvieux, M.C.; Peuchant, E.; Richard-Harston, S.; Decamps, A.; Reignier, B.; Emeriau, J.P.; Rainfray, M. Antioxidant defences and oxidative stress markers in erythrocytes and plasma from normally nourished elderly Alzheimer patients. Age. Ageing 2001, 30, 235–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Kaneko, S.; Kawakami, S.; Hara, Y.; Wakamori, M.; Itoh, E.; Minami, T.; Takada, Y.; Kume, T.; Katsuki, H.; Mori, Y.; et al. A critical role of TRPM2 in neuronal cell death by hydrogen peroxide. J. Pharm. Sci. 2006, 101, 66–76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Jin, S.M.; Cho, H.J.; Jung, E.S.; Shim, M.Y.; Mook-Jung, I. DNA damage-inducing agents elicit γ-secretase activation mediated by oxidative stress. Cell. Death Differ. 2008, 15, 1375–1384. [Google Scholar] [CrossRef] [PubMed]
  277. Muche, A.; Arendt, T.; Schliebs, R. Oxidative stress affects processing of amyloid precursor protein in vascular endothelial cells. PLoS ONE 2017, 12, e0178127. [Google Scholar] [CrossRef] [Green Version]
  278. Tamagno, E.; Bardini, P.; Obbili, A.; Vitali, A.; Borghi, R.; Zaccheo, D.; Pronzato, M.A.; Danni, O.; Smith, M.A.; Perry, G.; et al. Oxidative stress increases expression and activity of BACE in NT2 neurons. Neurobiol. Dis. 2002, 10, 279–288. [Google Scholar] [CrossRef] [Green Version]
  279. Tamagno, E.; Guglielmotto, M.; Aragno, M.; Borghi, R.; Autelli, R.; Giliberto, L.; Muraca, G.; Danni, O.; Zhu, X.; Smith, M.A.; et al. Oxidative stress activates a positive feedback between the gamma- and beta-secretase cleavages of the beta-amyloid precursor protein. J. Neurochem. 2008, 104, 683–695. [Google Scholar] [CrossRef] [Green Version]
  280. Jo, D.G.; Arumugam, T.V.; Woo, H.N.; Park, J.S.; Tang, S.C.; Mughal, M.; Hyun, D.H.; Park, J.H.; Choi, Y.H.; Gwon, A.R.; et al. Evidence that gamma-secretase mediates oxidative stress-induced beta-secretase expression in Alzheimer’s disease. Neurobiol. Aging 2010, 31, 917–925. [Google Scholar] [CrossRef] [Green Version]
  281. Hay, D.L.; Chen, S.; Lutz, T.A.; Parkes, D.G.; Roth, J.D. Amylin: Pharmacology, Physiology, and Clinical Potential. Pharmacol. Rev. 2015, 67, 564. [Google Scholar] [CrossRef] [Green Version]
  282. Ly, H.; Verma, N.; Sharma, S.; Kotiya, D.; Despa, S.; Abner, E.L.; Nelson, P.T.; Jicha, G.A.; Wilcock, D.M.; Goldstein, L.B.; et al. The association of circulating amylin with β-amyloid in familial Alzheimer’s disease. Alzheimers Dement. 2021, 7, e12130. [Google Scholar] [CrossRef] [PubMed]
  283. Ly, H.; Verma, N.; Kotiya, D.; Goldstein, L.B.; Jicha, G.A.; Despa, F. Amyloid-β removal from the brain is blocked by circulating amyloid-forming amylin secreted by the pancreas. Alzheimer’s Dement. 2021, 17, e053782. [Google Scholar] [CrossRef]
  284. Jhamandas, J.H.; Mactavish, D. β-Amyloid protein (Aβ) and human amylin regulation of apoptotic genes occurs through the amylin receptor. Apoptosis 2012, 17, 37–47. [Google Scholar] [CrossRef] [PubMed]
  285. Bharadwaj, P.; Solomon, T.; Sahoo, B.R.; Ignasiak, K.; Gaskin, S.; Rowles, J.; Verdile, G.; Howard, M.J.; Bond, C.S.; Ramamoorthy, A.; et al. Amylin and beta amyloid proteins interact to form amorphous heterocomplexes with enhanced toxicity in neuronal cells. Sci. Rep. 2020, 10, 10356. [Google Scholar] [CrossRef]
  286. Westergard, L.; Christensen, H.M.; Harris, D.A. The cellular prion protein (PrPC): Its physiological function and role in disease. Biochim. Biophys. Acta. (BBA.) Mol. Basis. Dis. 2007, 1772, 629–644. [Google Scholar] [CrossRef] [Green Version]
  287. Moser, M.; Colello, R.J.; Pott, U.; Oesch, B. Developmental expression of the prion protein gene in glial cells. Neuron 1995, 14, 509–517. [Google Scholar] [CrossRef] [Green Version]
  288. Zanusso, G.; Liu, D.; Ferrari, S.; Hegyi, I.; Yin, X.; Aguzzi, A.; Hornemann, S.; Liemann, S.; Glockshuber, R.; Manson, J.C.; et al. Prion protein expression in different species: Analysis with a panel of new mAbs. Proc. Natl. Acad. Sci. USA 1998, 95, 8812–8816. [Google Scholar] [CrossRef] [Green Version]
  289. Zhang, Y.; Sloan, S.A.; Clarke, L.E.; Caneda, C.; Plaza, C.A.; Blumenthal, P.D.; Vogel, H.; Steinberg, G.K.; Edwards, M.S.; Li, G.; et al. Purification and Characterization of Progenitor and Mature Human Astrocytes Reveals Transcriptional and Functional Differences with Mouse. Neuron 2016, 89, 37–53. [Google Scholar] [CrossRef] [Green Version]
  290. Gatti, J.-L.; Métayer, S.; Moudjou, M.; Andréoletti, O.; Lantier, F.d.r.; Dacheux, J.-L.; Sarradin, P. Prion Protein Is Secreted in Soluble Forms in the Epididymal Fluid and Proteolytically Processed and Transported in Seminal Plasma1. Biol. Reprod. 2002, 67, 393–400. [Google Scholar] [CrossRef] [Green Version]
  291. Tagliavini, F.; Prelli, F.; Porro, M.; Salmona, M.; Bugiani, O.; Frangione, B. A soluble form of prion protein in human cerebrospinal fluid: Implications for prion-related encephalopathies. Biochem. Biophys. Res. Commun. 1992, 184, 1398–1404. [Google Scholar] [CrossRef]
  292. Robinson, S.W.; Nugent, M.L.; Dinsdale, D.; Steinert, J.R. Prion protein facilitates synaptic vesicle release by enhancing release probability. Hum. Mol. Genet. 2014, 23, 4581–4596. [Google Scholar] [CrossRef] [PubMed]
  293. Collinge, J.; Whittington, M.A.; Sidle, K.C.; Smith, C.J.; Palmer, M.S.; Clarke, A.R.; Jefferys, J.G. Prion protein is necessary for normal synaptic function. Nature 1994, 370, 295–297. [Google Scholar] [CrossRef] [PubMed]
  294. Kanaani, J.; Prusiner, S.B.; Diacovo, J.; Baekkeskov, S.; Legname, G. Recombinant prion protein induces rapid polarization and development of synapses in embryonic rat hippocampal neurons in vitro. J. Neurochem. 2005, 95, 1373–1386. [Google Scholar] [CrossRef] [PubMed]
  295. Aguzzi, A.; Calella, A.M. Prions: Protein Aggregation and Infectious Diseases. Physiol. Rev. 2009, 89, 1105–1152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Ma, J.; Lindquist, S. Conversion of PrP to a Self-Perpetuating PrPSc-like Conformation in the Cytosol. Science 2002, 298, 1785–1788. [Google Scholar] [CrossRef] [PubMed]
  297. Ferrer, I.; Blanco, R.; Carmona, M.; Puig, B.; Ribera, R.; Rey, M.J.; Ribalta, T. Prion protein expression in senile plaques in Alzheimer’s disease. Acta Neuropathol. 2001, 101, 49–56. [Google Scholar] [CrossRef]
  298. Schwarze-Eicker, K.; Keyvani, K.; Görtz, N.; Westaway, D.; Sachser, N.; Paulus, W. Prion protein (PrPc) promotes β-amyloid plaque formation. Neurobiol. Aging 2005, 26, 1177–1182. [Google Scholar] [CrossRef]
  299. Takahashi, R.H.; Yokotsuka, M.; Tobiume, M.; Sato, Y.; Hasegawa, H.; Nagao, T.; Gouras, G.K. Accumulation of cellular prion protein within β-amyloid oligomer plaques in aged human brains. Brain. Pathol. 2021, 31, e12941. [Google Scholar] [CrossRef]
  300. Dohler, F.; Sepulveda-Falla, D.; Krasemann, S.; Altmeppen, H.; Schlüter, H.; Hildebrand, D.; Zerr, I.; Matschke, J.; Glatzel, M. High molecular mass assemblies of amyloid-β oligomers bind prion protein in patients with Alzheimer’s disease. Brain 2014, 137, 873–886. [Google Scholar] [CrossRef]
  301. Falker, C.; Hartmann, A.; Guett, I.; Dohler, F.; Altmeppen, H.; Betzel, C.; Schubert, R.; Thurm, D.; Wegwitz, F.; Joshi, P.; et al. Exosomal cellular prion protein drives fibrillization of amyloid beta and counteracts amyloid beta-mediated neurotoxicity. J. Neurochem. 2016, 137, 88–100. [Google Scholar] [CrossRef] [Green Version]
  302. Gimbel, D.A.; Nygaard, H.B.; Coffey, E.E.; Gunther, E.C.; Lauren, J.; Gimbel, Z.A.; Strittmatter, S.M. Memory Impairment in Transgenic Alzheimer Mice Requires Cellular Prion Protein. J. Neurosci. 2010, 30, 6367–6374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Salazar, S.V.; Gallardo, C.; Kaufman, A.C.; Herber, C.S.; Haas, L.T.; Robinson, S.; Manson, J.C.; Lee, M.K.; Strittmatter, S.M. Conditional Deletion of Prnp Rescues Behavioral and Synaptic Deficits after Disease Onset in Transgenic Alzheimer’s Disease. J. Neurosci. 2017, 37, 9207–9221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Parkin, E.T.; Watt, N.T.; Hussain, I.; Eckman, E.A.; Eckman, C.B.; Manson, J.C.; Baybutt, H.N.; Turner, A.J.; Hooper, N.M. Cellular prion protein regulates beta-secretase cleavage of the Alzheimer’s amyloid precursor protein. Proc. Natl. Acad. Sci. USA 2007, 104, 11062–11067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Griffiths, H.H.; Whitehouse, I.J.; Baybutt, H.; Brown, D.; Kellett, K.A.; Jackson, C.D.; Turner, A.J.; Piccardo, P.; Manson, J.C.; Hooper, N.M. Prion protein interacts with BACE1 protein and differentially regulates its activity toward wild type and Swedish mutant amyloid precursor protein. J. Biol. Chem. 2011, 286, 33489–33500. [Google Scholar] [CrossRef] [Green Version]
  306. Yeh, F.L.; Hansen, D.V.; Sheng, M. TREM2, Microglia, and Neurodegenerative Diseases. Trends Mol. Med. 2017, 23, 512–533. [Google Scholar] [CrossRef]
  307. Zhao, Y.; Wu, X.; Li, X.; Jiang, L.L.; Gui, X.; Liu, Y.; Sun, Y.; Zhu, B.; Pina-Crespo, J.C.; Zhang, M.; et al. TREM2 Is a Receptor for beta-Amyloid that Mediates Microglial Function. Neuron 2018, 97, 1023–1031.e7. [Google Scholar] [CrossRef] [Green Version]
  308. Parhizkar, S.; Arzberger, T.; Brendel, M.; Kleinberger, G.; Deussing, M.; Focke, C.; Nuscher, B.; Xiong, M.; Ghasemigharagoz, A.; Katzmarski, N.; et al. Loss of TREM2 function increases amyloid seeding but reduces plaque-associated ApoE. Nat. Neurosci. 2019, 22, 191–204. [Google Scholar] [CrossRef]
  309. Wang, Y.; Ulland, T.K.; Ulrich, J.D.; Song, W.; Tzaferis, J.A.; Hole, J.T.; Yuan, P.; Mahan, T.E.; Shi, Y.; Gilfillan, S.; et al. TREM2-mediated early microglial response limits diffusion and toxicity of amyloid plaques. J. Exp. Med. 2016, 213, 667–675. [Google Scholar] [CrossRef] [Green Version]
  310. Fu, W.; Vukojevic, V.; Patel, A.; Soudy, R.; MacTavish, D.; Westaway, D.; Kaur, K.; Goncharuk, V.; Jhamandas, J. Role of microglial amylin receptors in mediating beta amyloid (Abeta)-induced inflammation. J. Neuroinflammation. 2017, 14, 199. [Google Scholar] [CrossRef] [Green Version]
  311. Bailey, R.J.; Bradley, J.W.; Poyner, D.R.; Rathbone, D.L.; Hay, D.L. Functional characterization of two human receptor activity-modifying protein 3 variants. Peptides 2010, 31, 579–584. [Google Scholar] [CrossRef]
  312. Carrasquillo, M.M.; Belbin, O.; Hunter, T.A.; Ma, L.; Bisceglio, G.D.; Zou, F.; Crook, J.E.; Pankratz, V.S.; Sando, S.B.; Aasly, J.O.; et al. Replication of EPHA1 and CD33 associations with late-onset Alzheimer’s disease: A multi-centre case-control study. Mol. Neurodegener. 2011, 6, 54. [Google Scholar] [CrossRef] [PubMed]
  313. Hollingworth, P.; Harold, D.; Sims, R.; Gerrish, A.; Lambert, J.C.; Carrasquillo, M.M.; Abraham, R.; Hamshere, M.L.; Pahwa, J.S.; Moskvina, V.; et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s disease. Nat. Genet. 2011, 43, 429–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Naj, A.C.; Jun, G.; Beecham, G.W.; Wang, L.S.; Vardarajan, B.N.; Buros, J.; Gallins, P.J.; Buxbaum, J.D.; Jarvik, G.P.; Crane, P.K.; et al. Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are associated with late-onset Alzheimer’s disease. Nat. Genet. 2011, 43, 436–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Dos Santos, L.R.; Pimassoni, L.H.S.; Sena, G.G.S.; Camporez, D.; Belcavello, L.; Trancozo, M.; Morelato, R.L.; Errera, F.I.V.; Bueno, M.R.P.; de Paula, F. Validating GWAS Variants from Microglial Genes Implicated in Alzheimer’s Disease. J. Mol. Neurosci. 2017, 62, 215–221. [Google Scholar] [CrossRef]
  316. Zhou, L.; Li, H.Y.; Wang, J.H.; Deng, Z.Z.; Shan, Y.L.; Tan, S.; Shi, Y.H.; Zhang, M.X.; Liu, S.X.; Zhang, B.J.; et al. Correlation of gene polymorphisms of CD36 and ApoE with susceptibility of Alzheimer disease: A case-control study. Medicine 2018, 97, e12470. [Google Scholar] [CrossRef]
  317. Sery, O.; Janoutova, J.; Ewerlingova, L.; Halova, A.; Lochman, J.; Janout, V.; Khan, N.A.; Balcar, V.J. CD36 gene polymorphism is associated with Alzheimer’s disease. Biochimie 2017, 135, 46–53. [Google Scholar] [CrossRef]
  318. Peng, L.; Yu, Y.; Liu, J.; Li, S.; He, H.; Cheng, N.; Ye, R.D. The chemerin receptor CMKLR1 is a functional receptor for amyloid-beta peptide. J. Alzheimers Dis. 2015, 43, 227–242. [Google Scholar] [CrossRef]
  319. Brandenburg, L.O.; Konrad, M.; Wruck, C.J.; Koch, T.; Lucius, R.; Pufe, T. Functional and physical interactions between formyl-peptide-receptors and scavenger receptor MARCO and their involvement in amyloid beta 1-42-induced signal transduction in glial cells. J. Neurochem. 2010, 113, 749–760. [Google Scholar] [CrossRef]
  320. Zhang, D.; Hu, X.; Qian, L.; Chen, S.H.; Zhou, H.; Wilson, B.; Miller, D.S.; Hong, J.S. Microglial MAC1 receptor and PI3K are essential in mediating beta-amyloid peptide-induced microglial activation and subsequent neurotoxicity. J. Neuroinflammation 2011, 8, 3. [Google Scholar] [CrossRef] [Green Version]
  321. Roos, D.; Law, S.K. Hematologically important mutations: Leukocyte adhesion deficiency. Blood Cells Mol. Dis. 2001, 27, 1000–1004. [Google Scholar] [CrossRef]
  322. Alarcon, R.; Fuenzalida, C.; Santibanez, M.; von Bernhardi, R. Expression of scavenger receptors in glial cells. Comparing the adhesion of astrocytes and microglia from neonatal rats to surface-bound beta-amyloid. J. Biol. Chem. 2005, 280, 30406–30415. [Google Scholar] [CrossRef] [PubMed]
  323. Morkuniene, R.; Cizas, P.; Jankeviciute, S.; Petrolis, R.; Arandarcikaite, O.; Krisciukaitis, A.; Borutaite, V. Small Abeta1-42 oligomer-induced membrane depolarization of neuronal and microglial cells: Role of N-methyl-D-aspartate receptors. J. Neurosci. Res. 2015, 93, 475–486. [Google Scholar] [CrossRef] [PubMed]
  324. Jiang, H.; Jia, J. Association between NR2B subunit gene (GRIN2B) promoter polymorphisms and sporadic Alzheimer’s disease in the North Chinese population. Neurosci. Lett. 2009, 450, 356–360. [Google Scholar] [CrossRef] [PubMed]
  325. Liu, H.P.; Lin, W.Y.; Liu, S.H.; Wang, W.F.; Tsai, C.H.; Wu, B.T.; Wang, C.K.; Tsai, F.J. Genetic variation in N-methyl-D-aspartate receptor subunit NR3A but not NR3B influences susceptibility to Alzheimer’s disease. Dement. Geriatr. Cogn. Disord. 2009, 28, 521–527. [Google Scholar] [CrossRef] [PubMed]
  326. Andreoli, V.; De Marco, E.V.; Trecroci, F.; Cittadella, R.; Di Palma, G.; Gambardella, A. Potential involvement of GRIN2B encoding the NMDA receptor subunit NR2B in the spectrum of Alzheimer’s disease. J. Neural. Transm. 2014, 121, 533–542. [Google Scholar] [CrossRef]
  327. Stein, J.L.; Hua, X.; Morra, J.H.; Lee, S.; Hibar, D.P.; Ho, A.J.; Leow, A.D.; Toga, A.W.; Sul, J.H.; Kang, H.M.; et al. Genome-wide analysis reveals novel genes influencing temporal lobe structure with relevance to neurodegeneration in Alzheimer’s disease. Neuroimage 2010, 51, 542–554. [Google Scholar] [CrossRef] [Green Version]
  328. Chen, C.; Li, X.; Wang, T.; Wang, H.H.; Fu, Y.; Zhang, L.; Xiao, S.F. Association between NMDA receptor subunit 2b gene polymorphism and Alzheimer’s disease in Chinese Han population in Shanghai. Neurosci. Bull. 2010, 26, 395–400. [Google Scholar] [CrossRef] [Green Version]
  329. Tsai, S.J.; Liu, H.C.; Liu, T.Y.; Cheng, C.Y.; Hong, C.J. Association analysis for the genetic variants of the NMDA receptor subunit 2b and Alzheimer’s disease. Dement. Geriatr. Cogn. Disord. 2002, 13, 91–94. [Google Scholar] [CrossRef]
  330. Ozawa, D.; Nakamura, T.; Koike, M.; Hirano, K.; Miki, Y.; Beppu, M. Shuttling Protein Nucleolin Is a Microglia Receptor for Amyloid Beta Peptide 1-42. Biol. Pharm. Bull. 2013, 36, 1587–1593. [Google Scholar] [CrossRef] [Green Version]
  331. Lue, L.F.; Walker, D.G.; Brachova, L.; Beach, T.G.; Rogers, J.; Schmidt, A.M.; Stern, D.M.; Yan, S.D. Involvement of microglial receptor for advanced glycation endproducts (RAGE) in Alzheimer’s disease: Identification of a cellular activation mechanism. Exp. Neurol. 2001, 171, 29–45. [Google Scholar] [CrossRef]
  332. Li, K.; Dai, D.; Zhao, B.; Yao, L.; Yao, S.; Wang, B.; Yang, Z. Association between the RAGE G82S polymorphism and Alzheimer’s disease. J. Neural. Transm. 2010, 117, 97–104. [Google Scholar] [CrossRef] [PubMed]
  333. Daborg, J.; von Otter, M.; Sjolander, A.; Nilsson, S.; Minthon, L.; Gustafson, D.R.; Skoog, I.; Blennow, K.; Zetterberg, H. Association of the RAGE G82S polymorphism with Alzheimer’s disease. J. Neural. Transm. 2010, 117, 861–867. [Google Scholar] [CrossRef] [Green Version]
  334. Atac, Z.S.; Alaylioglu, M.; Dursun, E.; Gezen-Ak, D.; Yilmazer, S.; Gurvit, H. G82S polymorphism of receptor for advanced glycation end products gene and serum soluble RAGE levels in mild cognitive impairment and dementia of Alzheimer’s type patients in Turkish population. J. Clin. Neurosci. 2019, 59, 197–201. [Google Scholar] [CrossRef] [PubMed]
  335. Pericak-Vance, M.A.; Bass, M.P.; Yamaoka, L.H.; Gaskell, P.C.; Scott, W.K.; Terwedow, H.A.; Menold, M.M.; Conneally, P.M.; Small, G.W.; Vance, J.M.; et al. Complete genomic screen in late-onset familial Alzheimer disease. Evidence for a new locus on chromosome 12. JAMA 1997, 278, 1237–1241. [Google Scholar] [CrossRef] [PubMed]
  336. Luedecking-Zimmer, E.; DeKosky, S.T.; Nebes, R.; Kamboh, M.I. Association of the 3′ UTR transcription factor LBP-1c/CP2/LSF polymorphism with late-onset Alzheimer’s disease. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2003, 117B, 114–117. [Google Scholar] [CrossRef] [PubMed]
  337. Vergeer, M.; Korporaal, S.J.; Franssen, R.; Meurs, I.; Out, R.; Hovingh, G.K.; Hoekstra, M.; Sierts, J.A.; Dallinga-Thie, G.M.; Motazacker, M.M.; et al. Genetic variant of the scavenger receptor BI in humans. N. Engl. J. Med. 2011, 364, 136–145. [Google Scholar] [CrossRef]
  338. Zanoni, P.; Khetarpal, S.A.; Larach, D.B.; Hancock-Cerutti, W.F.; Millar, J.S.; Cuchel, M.; DerOhannessian, S.; Kontush, A.; Surendran, P.; Saleheen, D.; et al. Rare variant in scavenger receptor BI raises HDL cholesterol and increases risk of coronary heart disease. Science 2016, 351, 1166–1171. [Google Scholar] [CrossRef] [Green Version]
  339. Nakamura, K.; Ohya, W.; Funakoshi, H.; Sakaguchi, G.; Kato, A.; Takeda, M.; Kudo, T.; Nakamura, T. Possible role of scavenger receptor SRCL in the clearance of amyloid-beta in Alzheimer’s disease. J. Neurosci. Res. 2006, 84, 874–890. [Google Scholar] [CrossRef]
  340. Fassbender, K.; Walter, S.; Kuhl, S.; Landmann, R.; Ishii, K.; Bertsch, T.; Stalder, A.K.; Muehlhauser, F.; Liu, Y.; Ulmer, A.J.; et al. The LPS receptor (CD14) links innate immunity with Alzheimer’s disease. FASEB J. 2004, 18, 203–205. [Google Scholar] [CrossRef]
  341. Hamann, L.; Bustami, J.; Iakoubov, L.; Szwed, M.; Mossakowska, M.; Schumann, R.R.; Puzianowska-Kuznicka, M. TLR-6 SNP P249S is associated with healthy aging in nonsmoking Eastern European Caucasians—A cohort study. Immun. Ageing 2016, 13, 7. [Google Scholar] [CrossRef] [Green Version]
  342. Wang, Y.; Wu, X.; Deng, X.; Ma, Y.; Huang, S.; Wang, Y. Association of CD14-260 (-159) C/T and Alzheimer’s disease: Systematic review and trial sequential analyses. J. Neural. Transm. 2018, 125, 1313–1318. [Google Scholar] [CrossRef] [PubMed]
  343. Reed-Geaghan, E.G.; Savage, J.C.; Hise, A.G.; Landreth, G.E. CD14 and Toll-Like Receptors 2 and 4 Are Required for Fibrillar A beta-Stimulated Microglial Activation. J. Neurosci. 2009, 29, 11982–11992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Jiang, T.; Zhang, Y.D.; Gao, Q.; Zhou, J.S.; Zhu, X.C.; Lu, H.; Shi, J.Q.; Tan, L.; Chen, Q.; Yu, J.T. TREM1 facilitates microglial phagocytosis of amyloid beta. Acta Neuropathol. 2016, 132, 667–683. [Google Scholar] [CrossRef]
  345. Zhong, L.; Wang, Z.; Wang, D.; Wang, Z.; Martens, Y.A.; Wu, L.; Xu, Y.; Wang, K.; Li, J.; Huang, R.; et al. Amyloid-beta modulates microglial responses by binding to the triggering receptor expressed on myeloid cells 2 (TREM2). Mol. Neurodegener. 2018, 13, 15. [Google Scholar] [CrossRef]
  346. Yuan, P.; Condello, C.; Keene, C.D.; Wang, Y.; Bird, T.D.; Paul, S.M.; Luo, W.; Colonna, M.; Baddeley, D.; Grutzendler, J. TREM2 Haplodeficiency in Mice and Humans Impairs the Microglia Barrier Function Leading to Decreased Amyloid Compaction and Severe Axonal Dystrophy. Neuron 2016, 90, 724–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  347. Piccio, L.; Deming, Y.; Del-Aguila, J.L.; Ghezzi, L.; Holtzman, D.M.; Fagan, A.M.; Fenoglio, C.; Galimberti, D.; Borroni, B.; Cruchaga, C. Cerebrospinal fluid soluble TREM2 is higher in Alzheimer disease and associated with mutation status. Acta Neuropathol. 2016, 131, 925–933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  348. Guerreiro, R.; Wojtas, A.; Bras, J.; Carrasquillo, M.; Rogaeva, E.; Majounie, E.; Cruchaga, C.; Sassi, C.; Kauwe, J.S.K.; Lupton, M.K.; et al. TREM2 Variants in Alzheimer’s Disease. New Engl. J. Med. 2013, 368, 117–127. [Google Scholar] [CrossRef] [Green Version]
  349. Slattery, C.F.; Beck, J.A.; Harper, L.; Adamson, G.; Abdi, Z.; Uphill, J.; Campbell, T.; Druyeh, R.; Mahoney, C.J.; Rohrer, J.D.; et al. R47H TREM2 variant increases risk of typical early-onset Alzheimer’s disease but not of prion or frontotemporal dementia. Alzheimers Dement. 2014, 10, 602–608. [Google Scholar] [CrossRef] [Green Version]
  350. Kleinberger, G.; Yamanishi, Y.; Suarez-Calvet, M.; Czirr, E.; Lohmann, E.; Cuyvers, E.; Struyfs, H.; Pettkus, N.; Wenninger-Weinzierl, A.; Mazaheri, F.; et al. TREM2 mutations implicated in neurodegeneration impair cell surface transport and phagocytosis. Sci. Transl. Med. 2014, 6, 243ra286. [Google Scholar] [CrossRef]
  351. Benitez, B.A.; Cooper, B.; Pastor, P.; Jin, S.C.; Lorenzo, E.; Cervantes, S.; Cruchaga, C. TREM2 is associated with the risk of Alzheimer’s disease in Spanish population. Neurobiol. Aging 2013, 34. [Google Scholar] [CrossRef] [Green Version]
  352. Jiang, T.; Hou, J.K.; Gao, Q.; Yu, J.T.; Zhou, J.S.; Zhao, H.D.; Zhang, Y.D. TREM2 p.H157Y Variant and the Risk of Alzheimer’s Disease: A Meta-Analysis Involving 14,510 Subjects. Curr. Neurovasc. Res. 2016, 13, 318–320. [Google Scholar] [CrossRef] [PubMed]
  353. Li, R.; Wang, X.; He, P.F. The most prevalent rare coding variants of TREM2 conferring risk of Alzheimer’s disease: A systematic review and meta-analysis. Exp. Med. 2021, 21. [Google Scholar] [CrossRef] [PubMed]
  354. Koenigsknecht, J.; Landreth, G. Microglial phagocytosis of fibrillar beta-amyloid through a beta1 integrin-dependent mechanism. J. Neurosci. 2004, 24, 9838–9846. [Google Scholar] [CrossRef] [Green Version]
  355. Takeda, K.; Kaisho, T.; Akira, S. Toll-Like Receptors. Annu. Rev. Immunol. 2003, 21, 335–376. [Google Scholar] [CrossRef] [PubMed]
  356. Kielian, T. Toll-like receptors in central nervous system glial inflammation and homeostasis. J. Neurosci. Res. 2006, 83, 711–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  357. Fiebich, B.L.; Batista, C.R.A.; Saliba, S.W.; Yousif, N.M.; de Oliveira, A.C.P. Role of Microglia TLRs in Neurodegeneration. Front. Cell. Neurosci. 2018, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  358. Song, M.; Jin, J.; Lim, J.E.; Kou, J.; Pattanayak, A.; Rehman, J.A.; Kim, H.D.; Tahara, K.; Lalonde, R.; Fukuchi, K. TLR4 mutation reduces microglial activation, increases Abeta deposits and exacerbates cognitive deficits in a mouse model of Alzheimer’s disease. J. Neuroinflammation 2011, 8, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  359. Walter, S.; Letiembre, M.; Liu, Y.; Heine, H.; Penke, B.; Hao, W.; Bode, B.; Manietta, N.; Walter, J.; Schulz-Schüffer, W.; et al. Role of the Toll-Like Receptor 4 in Neuroinflammation in Alzheimer’s Disease. Cell. Physiol. Biochem. 2007, 20, 947–956. [Google Scholar] [CrossRef]
  360. Jin, J.J.; Kim, H.D.; Maxwell, J.A.; Li, L.; Fukuchi, K. Toll-like receptor 4-dependent upregulation of cytokines in a transgenic mouse model of Alzheimer’s disease. J. Neuroinflammation 2008, 5, 23. [Google Scholar] [CrossRef] [Green Version]
  361. McDonald, C.L.; Hennessy, E.; Rubio-Araiz, A.; Keogh, B.; McCormack, W.; McGuirk, P.; Reilly, M.; Lynch, M.A. Inhibiting TLR2 activation attenuates amyloid accumulation and glial activation in a mouse model of Alzheimer’s disease. Brain Behav. Immun. 2016, 58, 191–200. [Google Scholar] [CrossRef]
  362. Glatz, J.C.; Luiken, J.F.P. Dynamic role of the transmembrane glycoprotein CD36 (SR-B2) in cellular fatty acid uptake and utilization. J. Lipid. Res. 2018, 59, 1084–1093. [Google Scholar] [CrossRef] [PubMed]
  363. Ioghen, O.; Chitoiu, L.; Gherghiceanu, M.; Ceafalan, L.C.; Hinescu, M.E. CD36—A novel molecular target in the neurovascular unit. Eur. J. Neurosci. 2021, 53, 2500–2510. [Google Scholar] [CrossRef] [PubMed]
  364. Bamberger, M.E.; Harris, M.E.; McDonald, D.R.; Husemann, J.; Landreth, G.E. A Cell Surface Receptor Complex for Fibrillar β-Amyloid Mediates Microglial Activation. J. Neurosci. 2003, 23, 2665–2674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Stewart, C.R.; Stuart, L.M.; Wilkinson, K.; van Gils, J.M.; Deng, J.; Halle, A.; Rayner, K.J.; Boyer, L.; Zhong, R.; Frazier, W.A.; et al. CD36 ligands promote sterile inflammation through assembly of a Toll-like receptor 4 and 6 heterodimer. Nat. Immunol. 2010, 11, 155–161. [Google Scholar] [CrossRef] [Green Version]
  366. Ricciarelli, R.; D’Abramo, C.; Zingg, J.M.; Giliberto, L.; Markesbery, W.; Azzi, A.; Marinari, U.M.; Pronzato, M.A.; Tabaton, M. CD36 overexpression in human brain correlates with beta-amyloid deposition but not with Alzheimer’s disease. Free Radic. Biol. Med. 2004, 36, 1018–1024. [Google Scholar] [CrossRef]
  367. Egana-Gorrono, L.; Lopez-Diez, R.; Yepuri, G.; Ramirez, L.S.; Reverdatto, S.; Gugger, P.F.; Shekhtman, A.; Ramasamy, R.; Schmidt, A.M. Receptor for Advanced Glycation End Products (RAGE) and Mechanisms and Therapeutic Opportunities in Diabetes and Cardiovascular Disease: Insights From Human Subjects and Animal Models. Front. Cardiovasc. Med. 2020, 7, 37. [Google Scholar] [CrossRef]
  368. Schmidt, A.M.; Vianna, M.; Gerlach, M.; Brett, J.; Ryan, J.; Kao, J.; Esposito, C.; Hegarty, H.; Hurley, W.; Clauss, M. Isolation and characterization of two binding proteins for advanced glycosylation end products from bovine lung which are present on the endothelial cell surface. J. Biol. Chem. 1992, 267, 14987–14997. [Google Scholar] [CrossRef]
  369. Schlueter, C.; Hauke, S.; Flohr, A.M.; Rogalla, P.; Bullerdiek, J. Tissue-specific expression patterns of the RAGE receptor and its soluble forms—A result of regulated alternative splicing? Biochim. Biophys. Acta 2003, 1630, 1–6. [Google Scholar] [CrossRef]
  370. Sparvero, L.J.; Asafu-Adjei, D.; Kang, R.; Tang, D.; Amin, N.; Im, J.; Rutledge, R.; Lin, B.; Amoscato, A.A.; Zeh, H.J.; et al. RAGE (Receptor for Advanced Glycation Endproducts), RAGE Ligands, and their role in Cancer and Inflammation. J. Transl. Med. 2009, 7, 17. [Google Scholar] [CrossRef] [Green Version]
  371. Chellappa, R.C.; Palanisamy, R.; Swaminathan, K. RAGE Isoforms, its Ligands and their Role in Pathophysiology of Alzheimer’s Disease. Curr. Alzheimer Res. 2020, 17, 1262–1279. [Google Scholar] [CrossRef]
  372. Yan, S.D.; Chen, X.; Fu, J.; Chen, M.; Zhu, H.; Roher, A.; Slattery, T.; Zhao, L.; Nagashima, M.; Morser, J.; et al. RAGE and amyloid-beta peptide neurotoxicity in Alzheimer’s disease. Nature 1996, 382, 685–691. [Google Scholar] [CrossRef] [PubMed]
  373. Deane, R.; Wu, Z.; Sagare, A.; Davis, J.; Du Yan, S.; Hamm, K.; Xu, F.; Parisi, M.; LaRue, B.; Hu, H.W.; et al. LRP/amyloid beta-peptide interaction mediates differential brain efflux of Abeta isoforms. Neuron 2004, 43, 333–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Fang, F.; Yu, Q.; Arancio, O.; Chen, D.; Gore, S.S.; Yan, S.S.; Yan, S.F. RAGE mediates Abeta accumulation in a mouse model of Alzheimer’s disease via modulation of beta- and gamma-secretase activity. Hum. Mol. Genet. 2018, 27, 1002–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  375. Abedini, A.; Cao, P.; Plesner, A.; Zhang, J.; He, M.; Derk, J.; Patil, S.A.; Rosario, R.; Lonier, J.; Song, F.; et al. RAGE binds preamyloid IAPP intermediates and mediates pancreatic β cell proteotoxicity. J. Clin. Investig. 2018, 128, 682–698. [Google Scholar] [CrossRef] [Green Version]
  376. Fang, F.; Lue, L.F.; Yan, S.; Xu, H.; Luddy, J.S.; Chen, D.; Walker, D.G.; Stern, D.M.; Yan, S.; Schmidt, A.M.; et al. RAGE-dependent signaling in microglia contributes to neuroinflammation, Abeta accumulation, and impaired learning/memory in a mouse model of Alzheimer’s disease. FASEB J. 2010, 24, 1043–1055. [Google Scholar] [CrossRef] [Green Version]
  377. Gao, J.L.; Chen, H.; Filie, J.D.; Kozak, C.A.; Murphy, P.M. Differential expansion of the N-formylpeptide receptor gene cluster in human and mouse. Genomics 1998, 51, 270–276. [Google Scholar] [CrossRef]
  378. Migeotte, I.; Communi, D.; Parmentier, M. Formyl peptide receptors: A promiscuous subfamily of G protein-coupled receptors controlling immune responses. Cytokine Growth Factor Rev. 2006, 17, 501–519. [Google Scholar] [CrossRef]
  379. Cattaneo, F.; Guerra, G.; Ammendola, R. Expression and signaling of formyl-peptide receptors in the brain. Neurochem. Res. 2010, 35, 2018–2026. [Google Scholar] [CrossRef]
  380. Bloes, D.A.; Kretschmer, D.; Peschel, A. Enemy attraction: Bacterial agonists for leukocyte chemotaxis receptors. Nat. Rev. Microbiol. 2015, 13, 95–104. [Google Scholar] [CrossRef]
  381. Dahlgren, C.; Gabl, M.; Holdfeldt, A.; Winther, M.; Forsman, H. Basic characteristics of the neutrophil receptors that recognize formylated peptides, a danger-associated molecular pattern generated by bacteria and mitochondria. Biochem. Pharmacol. 2016, 114, 22–39. [Google Scholar] [CrossRef]
  382. Brandenburg, L.-O.; Koch, T.; Sievers, J.; Lucius, R. Internalization of PrP106–126 by the formyl-peptide-receptor-like-1 in glial cells. J. Neurochem. 2007, 101, 718–728. [Google Scholar] [CrossRef] [PubMed]
  383. Zhang, S.; Gong, H.; Ge, Y.; Ye, R.D. Biased allosteric modulation of formyl peptide receptor 2 leads to distinct receptor conformational states for pro- and anti-inflammatory signaling. Pharm. Res. 2020, 161, 105117. [Google Scholar] [CrossRef] [PubMed]
  384. Schröder, N.; Schaffrath, A.; Welter, J.A.; Putzka, T.; Griep, A.; Ziegler, P.; Brandt, E.; Samer, S.; Heneka, M.T.; Kaddatz, H.; et al. Inhibition of formyl peptide receptors improves the outcome in a mouse model of Alzheimer disease. J. Neuroinflammation 2020, 17, 131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  385. Arredouani, M.S.; Palecanda, A.; Koziel, H.; Huang, Y.C.; Imrich, A.; Sulahian, T.H.; Ning, Y.Y.; Yang, Z.; Pikkarainen, T.; Sankala, M.; et al. MARCO is the major binding receptor for unopsonized particles and bacteria on human alveolar macrophages. J. Immunol. 2005, 175, 6058–6064. [Google Scholar] [CrossRef] [Green Version]
  386. Van der Laan, L.J.W.; Döpp, E.A.; Haworth, R.; Pikkarainen, T.; Kangas, M.; Elomaa, O.; Dijkstra, C.D.; Gordon, S.; Tryggvason, K.; Kraal, G. Regulation and Functional Involvement of Macrophage Scavenger Receptor MARCO in Clearance of Bacteria In Vivo. J. Immunol. 1999, 162, 939. [Google Scholar]
  387. Thelen, T.; Hao, Y.; Medeiros, A.I.; Curtis, J.L.; Serezani, C.H.; Kobzik, L.; Harris, L.H.; Aronoff, D.M. The class A scavenger receptor, macrophage receptor with collagenous structure, is the major phagocytic receptor for Clostridium sordellii expressed by human decidual macrophages. J. Immunol. 2010, 185, 4328–4335. [Google Scholar] [CrossRef] [Green Version]
  388. Sankala, M.; Brannstrom, A.; Schulthess, T.; Bergmann, U.; Morgunova, E.; Engel, J.; Tryggvason, K.; Pikkarainen, T. Characterization of recombinant soluble macrophage scavenger receptor MARCO. J. Biol. Chem. 2002, 277, 33378–33385. [Google Scholar] [CrossRef] [Green Version]
  389. Mukhopadhyay, S.; Varin, A.; Chen, Y.; Liu, B.; Tryggvason, K.; Gordon, S. SR-A/MARCO-mediated ligand delivery enhances intracellular TLR and NLR function, but ligand scavenging from cell surface limits TLR4 response to pathogens. Blood 2011, 117, 1319–1328. [Google Scholar] [CrossRef]
  390. Yun, H.; Dumbell, R.; Hanna, K.; Bowen, J.; McLean, S.L.; Kantamneni, S.; Pors, K.; Wu, Q.F.; Helfer, G. The Chemerin-CMKLR1 Axis is Functionally important for Central Regulation of Energy Homeostasis. Front. Physiol. 2022, 13, 897105. [Google Scholar] [CrossRef]
  391. Arita, M.; Ohira, T.; Sun, Y.P.; Elangovan, S.; Chiang, N.; Serhan, C.N. Resolvin E1 selectively interacts with leukotriene B4 receptor BLT1 and ChemR23 to regulate inflammation. J. Immunol. 2007, 178, 3912–3917. [Google Scholar] [CrossRef] [Green Version]
  392. Zhang, H.; Lin, A.; Gong, P.; Chen, Y.; Ye, R.D.; Qian, F.; Zhang, Y.; Yu, Y. The Chemokine-like Receptor 1 Deficiency Improves Cognitive Deficits of AD Mice and Attenuates Tau Hyperphosphorylation via Regulating Tau Seeding. J. Neurosci. 2020, 40, 6991–7007. [Google Scholar] [CrossRef] [PubMed]
  393. Massimino, M.L.; Simonato, M.; Spolaore, B.; Franchin, C.; Arrigoni, G.; Marin, O.; Monturiol-Gross, L.; Fernandez, J.; Lomonte, B.; Tonello, F. Cell surface nucleolin interacts with and internalizes Bothrops asper Lys49 phospholipase A2 and mediates its toxic activity. Sci. Rep. 2018, 8, 10619. [Google Scholar] [CrossRef] [PubMed]
  394. Tajrishi, M.M.; Tuteja, R.; Tuteja, N. Nucleolin: The most abundant multifunctional phosphoprotein of nucleolus. Commun. Integr. Biol. 2011, 4, 267–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Zaidi, S.H.E.; Malter, J.S. Nucleolin and Heterogeneous Nuclear Ribonucleoprotein C Proteins Specifically Interact with the 3′-Untranslated Region of Amyloid Protein Precursor mRNA. J. Biol. Chem. 1995, 270, 17292–17298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  396. Westmark, C.J.; Malter, J.S. The regulation of AβPP expression by RNA-binding proteins. Ageing Res. Rev. 2012, 11, 450–459. [Google Scholar] [CrossRef] [Green Version]
  397. Rajagopalan, L.E.; Westmark, C.J.; Jarzembowski, J.A.; Malter, J.S. hnRNP C increases amyloid precursor protein (APP) production by stabilizing APP mRNA. Nucleic Acids Res. 1998, 26, 3418–3423. [Google Scholar] [CrossRef] [Green Version]
  398. Belrose, J.C.; Jackson, M.F. TRPM2: A candidate therapeutic target for treating neurological diseases. Acta Pharm. Sin. 2018, 39, 722–732. [Google Scholar] [CrossRef] [Green Version]
  399. Tan, C.-H.; McNaughton, P.A. The TRPM2 ion channel is required for sensitivity to warmth. Nature 2016, 536, 460–463. [Google Scholar] [CrossRef] [Green Version]
  400. Malko, P.; Syed Mortadza, S.A.; McWilliam, J.; Jiang, L.H. TRPM2 Channel in Microglia as a New Player in Neuroinflammation Associated With a Spectrum of Central Nervous System Pathologies. Front. Pharm. 2019, 10, 239. [Google Scholar] [CrossRef] [Green Version]
  401. Fonfria, E.; Marshall, I.C.; Benham, C.D.; Boyfield, I.; Brown, J.D.; Hill, K.; Hughes, J.P.; Skaper, S.D.; McNulty, S. TRPM2 channel opening in response to oxidative stress is dependent on activation of poly(ADP-ribose) polymerase. Br. J. Pharm. 2004, 143, 186–192. [Google Scholar] [CrossRef] [Green Version]
  402. Aminzadeh, M.; Roghani, M.; Sarfallah, A.; Riazi, G.H. TRPM2 dependence of ROS-induced NLRP3 activation in Alzheimer’s disease. Int. Immunopharmacol. 2018, 54, 78–85. [Google Scholar] [CrossRef] [PubMed]
  403. Alawieyah Syed Mortadza, S.; Sim, J.A.; Neubrand, V.E.; Jiang, L.H. A critical role of TRPM2 channel in Abeta42 -induced microglial activation and generation of tumor necrosis factor-alpha. Glia 2018, 66, 562–575. [Google Scholar] [CrossRef] [PubMed]
  404. Ostapchenko, V.G.; Chen, M.; Guzman, M.S.; Xie, Y.F.; Lavine, N.; Fan, J.; Beraldo, F.H.; Martyn, A.C.; Belrose, J.C.; Mori, Y.; et al. The Transient Receptor Potential Melastatin 2 (TRPM2) Channel Contributes to beta-Amyloid Oligomer-Related Neurotoxicity and Memory Impairment. J. Neurosci. 2015, 35, 15157–15169. [Google Scholar] [CrossRef] [Green Version]
  405. Jiang, T.; Yu, J.T.; Hu, N.; Tan, M.S.; Zhu, X.C.; Tan, L. CD33 in Alzheimer’s disease. Mol. Neurobiol. 2014, 49, 529–535. [Google Scholar] [CrossRef]
  406. Zhao, L. CD33 in Alzheimer’s Disease—Biology, Pathogenesis, and Therapeutics: A Mini-Review. Gerontology 2019, 65, 323–331. [Google Scholar] [CrossRef] [PubMed]
  407. Wißfeld, J.; Nozaki, I.; Mathews, M.; Raschka, T.; Ebeling, C.; Hornung, V.; Brüstle, O.; Neumann, H. Deletion of Alzheimer’s disease-associated CD33 results in an inflammatory human microglia phenotype. Glia 2021, 69, 1393–1412. [Google Scholar] [CrossRef]
  408. Griciuc, A.; Serrano-Pozo, A.; Parrado, A.R.; Lesinski, A.N.; Asselin, C.N.; Mullin, K.; Hooli, B.; Choi, S.H.; Hyman, B.T.; Tanzi, R.E. Alzheimer’s disease risk gene CD33 inhibits microglial uptake of amyloid beta. Neuron 2013, 78, 631–643. [Google Scholar] [CrossRef] [Green Version]
  409. Griciuc, A.; Federico, A.N.; Natasan, J.; Forte, A.M.; McGinty, D.; Nguyen, H.; Volak, A.; LeRoy, S.; Gandhi, S.; Lerner, E.P.; et al. Gene therapy for Alzheimer’s disease targeting CD33 reduces amyloid beta accumulation and neuroinflammation. Hum. Mol. Genet. 2020, 29, 2920–2935. [Google Scholar] [CrossRef]
  410. Peress, N.S.; Fleit, H.B.; Perillo, E.; Kuljis, R.; Pezzullo, C. Identification of Fc gamma RI, II and III on normal human brain ramified microglia and on microglia in senile plaques in Alzheimer’s disease. J. Neuroimmunol. 1993, 48, 71–79. [Google Scholar] [CrossRef]
  411. Janus, C.; Pearson, J.; McLaurin, J.; Mathews, P.M.; Jiang, Y.; Schmidt, S.D.; Chishti, M.A.; Horne, P.; Heslin, D.; French, J.; et al. A beta peptide immunization reduces behavioural impairment and plaques in a model of Alzheimer’s disease. Nature 2000, 408, 979–982. [Google Scholar] [CrossRef]
  412. Bard, F.; Cannon, C.; Barbour, R.; Burke, R.L.; Games, D.; Grajeda, H.; Guido, T.; Hu, K.; Huang, J.; Johnson-Wood, K.; et al. Peripherally administered antibodies against amyloid beta-peptide enter the central nervous system and reduce pathology in a mouse model of Alzheimer disease. Nat. Med. 2000, 6, 916–919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  413. Söllvander, S.; Ekholm-Pettersson, F.; Brundin, R.M.; Westman, G.; Kilander, L.; Paulie, S.; Lannfelt, L.; Sehlin, D. Increased Number of Plasma B Cells Producing Autoantibodies Against Aβ42 Protofibrils in Alzheimer’s Disease. J. Alzheimers Dis. 2015, 48, 63–72. [Google Scholar] [CrossRef] [PubMed]
  414. Bulati, M.; Buffa, S.; Martorana, A.; Gervasi, F.; Camarda, C.; Azzarello, D.M.; Monastero, R.; Caruso, C.; Colonna-Romano, G. Double negative (IgG+IgD-CD27-) B cells are increased in a cohort of moderate-severe Alzheimer’s disease patients and show a pro-inflammatory trafficking receptor phenotype. J. Alzheimers Dis. 2015, 44, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  415. Bacskai, B.J.; Kajdasz, S.T.; McLellan, M.E.; Games, D.; Seubert, P.; Schenk, D.; Hyman, B.T. Non-Fc-mediated mechanisms are involved in clearance of amyloid-beta in vivo by immunotherapy. J. Neurosci. 2002, 22, 7873–7878. [Google Scholar] [CrossRef] [Green Version]
  416. Hardy, J.; Escott-Price, V. Genes, pathways and risk prediction in Alzheimer’s disease. Hum. Mol. Genet. 2019, 28, R235–R240. [Google Scholar] [CrossRef] [PubMed]
  417. Li, M.; Geng, R.; Li, C.; Meng, F.; Zhao, H.; Liu, J.; Dai, J.; Wang, X. Dysregulated gene-associated biomarkers for Alzheimer’s disease and aging. Transl. Neurosci. 2021, 12, 83–95. [Google Scholar] [CrossRef]
  418. Feuerbach, D.; Schindler, P.; Barske, C.; Joller, S.; Beng-Louka, E.; Worringer, K.A.; Kommineni, S.; Kaykas, A.; Ho, D.J.; Ye, C.; et al. ADAM17 is the main sheddase for the generation of human triggering receptor expressed in myeloid cells (hTREM2) ectodomain and cleaves TREM2 after Histidine 157. Neurosci. Lett. 2017, 660, 109–114. [Google Scholar] [CrossRef]
  419. Schlepckow, K.; Kleinberger, G.; Fukumori, A.; Feederle, R.; Lichtenthaler, S.F.; Steiner, H.; Haass, C. An Alzheimer-associated TREM2 variant occurs at the ADAM cleavage site and affects shedding and phagocytic function. EMBO. Mol. Med. 2017, 9, 1356–1365. [Google Scholar] [CrossRef]
  420. Lichtenthaler, S.F.; Tschirner, S.K.; Steiner, H. Secretases in Alzheimer’s disease: Novel insights into proteolysis of APP and TREM2. Curr. Opin. Neurobiol. 2022, 72, 101–110. [Google Scholar] [CrossRef]
  421. Zheng, H.; Jia, L.; Liu, C.C.; Rong, Z.; Zhong, L.; Yang, L.; Chen, X.F.; Fryer, J.D.; Wang, X.; Zhang, Y.W.; et al. TREM2 Promotes Microglial Survival by Activating Wnt/beta-Catenin Pathway. J. Neurosci. 2017, 37, 1772–1784. [Google Scholar] [CrossRef] [Green Version]
  422. Zhong, L.; Chen, X.F.; Wang, T.; Wang, Z.; Liao, C.; Wang, Z.; Huang, R.; Wang, D.; Li, X.; Wu, L.; et al. Soluble TREM2 induces inflammatory responses and enhances microglial survival. J. Exp. Med. 2017, 214, 597–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  423. Franzmeier, N.; Suarez-Calvet, M.; Frontzkowski, L.; Moore, A.; Hohman, T.J.; Morenas-Rodriguez, E.; Nuscher, B.; Shaw, L.; Trojanowski, J.Q.; Dichgans, M.; et al. Higher CSF sTREM2 attenuates ApoE4-related risk for cognitive decline and neurodegeneration. Mol. Neurodegener. 2020, 15, 57. [Google Scholar] [CrossRef] [PubMed]
  424. Thornton, P.; Sevalle, J.; Deery, M.J.; Fraser, G.; Zhou, Y.; Stahl, S.; Franssen, E.H.; Dodd, R.B.; Qamar, S.; Gomez Perez-Nievas, B.; et al. TREM2 shedding by cleavage at the H157-S158 bond is accelerated for the Alzheimer’s disease-associated H157Y variant. EMBO Mol. Med. 2017, 9, 1366–1378. [Google Scholar] [CrossRef] [PubMed]
  425. Zeng, F.; Shen, C.; Liu, Y.H.; Li, J.; Zhu, J.; Wang, Y.R.; Yan, J.C.; Gao, C.Y.; Zhou, H.D.; Deng, J.; et al. Genetic Association Between APP, ADAM10 Gene Polymorphism, and Sporadic Alzheimer’s Disease in the Chinese Population. Neurotox. Res. 2015, 27, 284–291. [Google Scholar] [CrossRef] [PubMed]
  426. Huang, W.H.; Chen, W.; Jiang, L.Y.; Yang, Y.X.; Yao, L.F.; Li, K.S. Influence of ADAM10 Polymorphisms on Plasma Level of Soluble Receptor for Advanced Glycation End Products and The Association With Alzheimer’s Disease Risk. Front. Genet. 2018, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  427. Criscuolo, C.; Fontebasso, V.; Middei, S.; Stazi, M.; Ammassari-Teule, M.; Yan, S.S.; Origlia, N. Entorhinal Cortex dysfunction can be rescued by inhibition of microglial RAGE in an Alzheimer’s disease mouse model. Sci. Rep. 2017, 7, 42370. [Google Scholar] [CrossRef] [Green Version]
  428. Zhang, L.; Bukulin, M.; Kojro, E.; Roth, A.; Metz, V.V.; Fahrenholz, F.; Nawroth, P.P.; Bierhaus, A.; Postina, R. Receptor for advanced glycation end products is subjected to protein ectodomain shedding by metalloproteinases. J. Biol. Chem. 2008, 283, 35507–35516. [Google Scholar] [CrossRef] [Green Version]
  429. Jang, Y.; Kim, J.Y.; Kang, S.M.; Kim, J.S.; Chae, J.S.; Kim, O.Y.; Koh, S.J.; Lee, H.C.; Ahn, C.W.; Song, Y.D.; et al. Association of the Gly82Ser polymorphism in the receptor for advanced glycation end products (RAGE) gene with circulating levels of soluble RAGE and inflammatory markers in nondiabetic and nonobese Koreans. Metabolism 2007, 56, 199–205. [Google Scholar] [CrossRef]
  430. Cohn, W.; Melnik, M.; Huang, C.; Teter, B.; Chandra, S.; Zhu, C.; McIntire, L.B.; John, V.; Gylys, K.H.; Bilousova, T. Multi-Omics Analysis of Microglial Extracellular Vesicles From Human Alzheimer’s Disease Brain Tissue Reveals Disease-Associated Signatures. Front. Pharm. 2021, 12, 766082. [Google Scholar] [CrossRef]
  431. Yang, H.; Wang, Y.; Kar, S. Effects of cholesterol transport inhibitor U18666A on APP metabolism in rat primary astrocytes. Glia. 2017, 65, 1728–1743. [Google Scholar] [CrossRef]
  432. Stavropoulou, A.V.; Mavrofrydi, O.; Saftig, P.; Efthimiopoulos, S. Serum Starvation Induces BACE1 Processing and Secretion. Curr. Alzheimer Res. 2017, 14, 453–459. [Google Scholar] [CrossRef] [PubMed]
  433. Nugent, A.A.; Lin, K.; van Lengerich, B.; Lianoglou, S.; Przybyla, L.; Davis, S.S.; Llapashtica, C.; Wang, J.; Kim, D.J.; Xia, D.; et al. TREM2 Regulates Microglial Cholesterol Metabolism upon Chronic Phagocytic Challenge. Neuron 2020, 105, 837–854 e839. [Google Scholar] [CrossRef] [PubMed]
  434. Erten-Lyons, D.; Woltjer, R.L.; Dodge, H.; Nixon, R.; Vorobik, R.; Calvert, J.F.; Leahy, M.; Montine, T.; Kaye, J. Factors associated with resistance to dementia despite high Alzheimer disease pathology. Neurology 2009, 72, 354–360. [Google Scholar] [CrossRef] [PubMed]
  435. Sloane, J.A.; Pietropaolo, M.F.; Rosene, D.L.; Moss, M.B.; Peters, A.; Kemper, T.; Abraham, C.R. Lack of correlation between plaque burden and cognition in the aged monkey. Acta. Neuropathol. 1997, 94, 471–478. [Google Scholar] [CrossRef] [PubMed]
  436. Petersen, R.C.; Aisen, P.; Boeve, B.F.; Geda, Y.E.; Ivnik, R.J.; Knopman, D.S.; Mielke, M.; Pankratz, V.S.; Roberts, R.; Rocca, W.A.; et al. Mild cognitive impairment due to Alzheimer disease in the community. Ann. Neurol. 2013, 74, 199–208. [Google Scholar] [CrossRef] [Green Version]
  437. Hsiao, K.; Chapman, P.; Nilsen, S.; Eckman, C.; Harigaya, Y.; Younkin, S.; Yang, F.; Cole, G. Correlative memory deficits, Abeta elevation, and amyloid plaques in transgenic mice. Science 1996, 274, 99–102. [Google Scholar] [CrossRef]
  438. Klyubin, I.; Betts, V.; Welzel, A.T.; Blennow, K.; Zetterberg, H.; Wallin, A.; Lemere, C.A.; Cullen, W.K.; Peng, Y.; Wisniewski, T.; et al. Amyloid beta protein dimer-containing human CSF disrupts synaptic plasticity: Prevention by systemic passive immunization. J. Neurosci. 2008, 28, 4231–4237. [Google Scholar] [CrossRef] [Green Version]
  439. Shankar, G.M.; Li, S.; Mehta, T.H.; Garcia-Munoz, A.; Shepardson, N.E.; Smith, I.; Brett, F.M.; Farrell, M.A.; Rowan, M.J.; Lemere, C.A.; et al. Amyloid-beta protein dimers isolated directly from Alzheimer’s brains impair synaptic plasticity and memory. Nat. Med. 2008, 14, 837–842. [Google Scholar] [CrossRef] [Green Version]
  440. Tomiyama, T.; Shimada, H. APP Osaka Mutation in Familial Alzheimer’s Disease-Its Discovery, Phenotypes, and Mechanism of Recessive Inheritance. Int. J. Mol. Sci. 2020, 21, 1413. [Google Scholar] [CrossRef] [Green Version]
  441. Faucher, P.; Mons, N.; Micheau, J.; Louis, C.; Beracochea, D.J. Hippocampal Injections of Oligomeric Amyloid β-peptide (1–42) Induce Selective Working Memory Deficits and Long-lasting Alterations of ERK Signaling Pathway. Front. Aging Neurosci. 2016, 7. [Google Scholar] [CrossRef] [Green Version]
  442. Balducci, C.; Beeg, M.; Stravalaci, M.; Bastone, A.; Sclip, A.; Biasini, E.; Tapella, L.; Colombo, L.; Manzoni, C.; Borsello, T.; et al. Synthetic amyloid-β oligomers impair long-term memory independently of cellular prion protein. Proc. Natl. Acad. Sci. USA 2010, 107, 2295–2300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  443. Cleary, J.P.; Walsh, D.M.; Hofmeister, J.J.; Shankar, G.M.; Kuskowski, M.A.; Selkoe, D.J.; Ashe, K.H. Natural oligomers of the amyloid-β protein specifically disrupt cognitive function. Nat. Neurosci. 2005, 8, 79–84. [Google Scholar] [CrossRef] [PubMed]
  444. Lesné, S.; Koh, M.T.; Kotilinek, L.; Kayed, R.; Glabe, C.G.; Yang, A.; Gallagher, M.; Ashe, K.H. A specific amyloid-β protein assembly in the brain impairs memory. Nature 2006, 440, 352–357. [Google Scholar] [CrossRef] [PubMed]
  445. Piller, C. Blots on a field? Science 2022, 377, 358–363. [Google Scholar] [CrossRef] [PubMed]
  446. Brinkmalm, G.; Hong, W.; Wang, Z.; Liu, W.; O’Malley, T.T.; Sun, X.; Frosch, M.P.; Selkoe, D.J.; Portelius, E.; Zetterberg, H.; et al. Identification of neurotoxic cross-linked amyloid-β dimers in the Alzheimer’s brain. Brain 2019, 142, 1441–1457. [Google Scholar] [CrossRef]
  447. Winkler, J.; Connor, D.J.; Frautschy, S.A.; Behl, C.; Waite, J.J.; Cole, G.M.; Thal, L.J. Lack of long-term effects after β-amyloid protein injections in rat brain. Neurobiol. Aging 1994, 15, 601–607. [Google Scholar] [CrossRef]
  448. Clemens, J.A.; Stephenson, D.T. Implants containing β-amyloid protein are not neurotoxic to young and old rat brain. Neurobiol. Aging 1992, 13, 581–586. [Google Scholar] [CrossRef]
  449. Benilova, I.; Karran, E.; De Strooper, B. The toxic Aβ oligomer and Alzheimer’s disease: An emperor in need of clothes. Nat. Neurosci. 2012, 15, 349–357. [Google Scholar] [CrossRef]
  450. Dahlgren, K.N.; Manelli, A.M.; Stine, W.B., Jr.; Baker, L.K.; Krafft, G.A.; LaDu, M.J. Oligomeric and fibrillar species of amyloid-beta peptides differentially affect neuronal viability. J. Biol. Chem. 2002, 277, 32046–32053. [Google Scholar] [CrossRef] [Green Version]
  451. Sgourakis, N.G.; Yan, Y.; McCallum, S.A.; Wang, C.; Garcia, A.E. The Alzheimer’s peptides Abeta40 and 42 adopt distinct conformations in water: A combined MD/NMR study. J. Mol. Biol. 2007, 368, 1448–1457. [Google Scholar] [CrossRef] [Green Version]
  452. Vivekanandan, S.; Brender, J.R.; Lee, S.Y.; Ramamoorthy, A. A partially folded structure of amyloid-beta(1-40) in an aqueous environment. Biochem. Biophys. Res. Commun. 2011, 411, 312–316. [Google Scholar] [CrossRef] [PubMed]
  453. Balbach, J.J.; Ishii, Y.; Antzutkin, O.N.; Leapman, R.D.; Rizzo, N.W.; Dyda, F.; Reed, J.; Tycko, R. Amyloid fibril formation by A beta 16-22, a seven-residue fragment of the Alzheimer’s beta-amyloid peptide, and structural characterization by solid state NMR. Biochemistry 2000, 39, 13748–13759. [Google Scholar] [CrossRef] [PubMed]
  454. Jahn, T.R.; Makin, O.S.; Morris, K.L.; Marshall, K.E.; Tian, P.; Sikorski, P.; Serpell, L.C. The common architecture of cross-beta amyloid. J. Mol. Biol. 2010, 395, 717–727. [Google Scholar] [CrossRef] [PubMed]
  455. Nelson, R.; Sawaya, M.R.; Balbirnie, M.; Madsen, A.Ø.; Riekel, C.; Grothe, R.; Eisenberg, D. Structure of the cross-β spine of amyloid-like fibrils. Nature 2005, 435, 773–778. [Google Scholar] [CrossRef] [Green Version]
  456. Faller, P.; Hureau, C. Reproducibility Problems of Amyloid-β Self-Assembly and How to Deal With Them. Front. Chem. 2021, 8, 611227. [Google Scholar] [CrossRef]
  457. Kheterpal, I.; Lashuel, H.A.; Hartley, D.M.; Walz, T.; Lansbury, P.T., Jr.; Wetzel, R. Abeta protofibrils possess a stable core structure resistant to hydrogen exchange. Biochemistry 2003, 42, 14092–14098. [Google Scholar] [CrossRef] [Green Version]
  458. Knowles, T.P.; Vendruscolo, M.; Dobson, C.M. The amyloid state and its association with protein misfolding diseases. Nat. Rev. Mol. Cell. Biol. 2014, 15, 384–396. [Google Scholar] [CrossRef]
  459. Armen, R.S.; DeMarco, M.L.; Alonso, D.O.; Daggett, V. Pauling and Corey’s alpha-pleated sheet structure may define the prefibrillar amyloidogenic intermediate in amyloid disease. Proc. Natl. Acad. Sci. USA 2004, 101, 11622–11627. [Google Scholar] [CrossRef] [Green Version]
  460. Shea, D.; Hsu, C.-C.; Bi, T.M.; Paranjapye, N.; Childers, M.C.; Cochran, J.; Tomberlin, C.P.; Wang, L.; Paris, D.; Zonderman, J.; et al. α-Sheet secondary structure in amyloid β-peptide drives aggregation and toxicity in Alzheimer’s disease. Proc. Natl. Acad. Sci. USA 2019, 116, 8895–8900. [Google Scholar] [CrossRef] [Green Version]
  461. Bi, T.M.; Daggett, V. The Role of α-sheet in Amyloid Oligomer Aggregation and Toxicity. Yale J. Biol. Med. 2018, 91, 247–255. [Google Scholar]
  462. Kasim, J.K.; Kavianinia, I.; Harris, P.W.R.; Brimble, M.A. Three Decades of Amyloid Beta Synthesis: Challenges and Advances. Front. Chem. 2019, 7, 472. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  463. Barrow, C.J.; Yasuda, A.; Kenny, P.T.; Zagorski, M.G. Solution conformations and aggregational properties of synthetic amyloid beta-peptides of Alzheimer’s disease. Analysis of circular dichroism spectra. J. Mol. Biol. 1992, 225, 1075–1093. [Google Scholar] [CrossRef]
  464. Jao, S.-C.; Ma, K.; Talafous, J.; Orlando, R.; Zagorski, M.G. Trifluoroacetic acid pretreatment reproducibly disaggregates the amyloid β-peptide. Amyloid 1997, 4, 240–252. [Google Scholar] [CrossRef]
  465. Foley, A.R.; Raskatov, J.A. Assessing Reproducibility in Amyloid β Research: Impact of Aβ Sources on Experimental Outcomes. ChemBioChem 2020, 21, 2425–2430. [Google Scholar] [CrossRef] [PubMed]
  466. Stine, W.B.; Jungbauer, L.; Yu, C.; LaDu, M.J. Preparing synthetic Aβ in different aggregation states. Methods Mol. Biol. 2011, 670, 13–32. [Google Scholar] [CrossRef] [Green Version]
  467. Ryan, T.M.; Caine, J.; Mertens, H.D.; Kirby, N.; Nigro, J.; Breheney, K.; Waddington, L.J.; Streltsov, V.A.; Curtain, C.; Masters, C.L.; et al. Ammonium hydroxide treatment of Aβ produces an aggregate free solution suitable for biophysical and cell culture characterization. PeerJ 2013, 1, e73. [Google Scholar] [CrossRef] [Green Version]
  468. Cruz, L.; Urbanc, B.; Borreguero, J.M.; Lazo, N.D.; Teplow, D.B.; Stanley, H.E. Solvent and mutation effects on the nucleation of amyloid β-protein folding. Proc. Natl. Acad. Sci. USA 2005, 102, 18258–18263. [Google Scholar] [CrossRef] [Green Version]
  469. Wei, G.; Shea, J.-E. Effects of Solvent on the Structure of the Alzheimer Amyloid-β(25–35) Peptide. Biophys. J. 2006, 91, 1638–1647. [Google Scholar] [CrossRef] [Green Version]
  470. Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
Figure 1. Membrane topology and domain architecture of microglial Aβ-binding receptors. Some receptors are shown exemplarily to underpin the diverging domain architecture und potential Aβ-binding site positions as far as identified (red arrows). Relative location of polymorphisms to binding sites are indicated by yellow diamonds. Proportion of domains has been neglected. For RAGE, binding of Aβ is assumed to occur as a dimer on the dimeric receptor [16]. In addition to cleaved soluble RAGE and TREM2, additional soluble splice forms exist [17,18].
Figure 1. Membrane topology and domain architecture of microglial Aβ-binding receptors. Some receptors are shown exemplarily to underpin the diverging domain architecture und potential Aβ-binding site positions as far as identified (red arrows). Relative location of polymorphisms to binding sites are indicated by yellow diamonds. Proportion of domains has been neglected. For RAGE, binding of Aβ is assumed to occur as a dimer on the dimeric receptor [16]. In addition to cleaved soluble RAGE and TREM2, additional soluble splice forms exist [17,18].
Cells 11 03421 g001
Figure 3. The amyloidogenic and non-amyloidogenic pathway of APP processing. (A) The non-amyloidogenic pathway is starting by α-secretase cleavage of APP which leads to the release of soluble APPα (sAPPα) and concomitant generation of the α-C-terminal fragment (α-CTF). Subsequent cleavage of the α-CTF by γ-secretase leads to the secretion of the small fragment p3 and release of the APP intracellular C-terminal domain (AICD) into the cytosol. In the amyloidogenic pathway, APP is first cleaved by β-secretase to produce soluble APPβ (sAPPβ) and the membrane retained β-C-terminal fragment (β-CTF). Further processing of the β-CTF by γ-secretase releases AICD as well as the Aβ peptide, which can form neurotoxic oligomers. Aβ production was reported to be higher in neuronal cultures compared to astrocytes or microglia cultures [95,96]. 3D structures (PDB: 1IYT) are based on NMR experiments by Crescenzi and colleagues [97]. (B) Glial cells mainly produce N-abridged Aβ peptides (up to 60%) such as Aβ2/3-X or Aβ4/5-x, while neurons produce predominantly Aβ peptides starting at position 1 (80%) [98].
Figure 3. The amyloidogenic and non-amyloidogenic pathway of APP processing. (A) The non-amyloidogenic pathway is starting by α-secretase cleavage of APP which leads to the release of soluble APPα (sAPPα) and concomitant generation of the α-C-terminal fragment (α-CTF). Subsequent cleavage of the α-CTF by γ-secretase leads to the secretion of the small fragment p3 and release of the APP intracellular C-terminal domain (AICD) into the cytosol. In the amyloidogenic pathway, APP is first cleaved by β-secretase to produce soluble APPβ (sAPPβ) and the membrane retained β-C-terminal fragment (β-CTF). Further processing of the β-CTF by γ-secretase releases AICD as well as the Aβ peptide, which can form neurotoxic oligomers. Aβ production was reported to be higher in neuronal cultures compared to astrocytes or microglia cultures [95,96]. 3D structures (PDB: 1IYT) are based on NMR experiments by Crescenzi and colleagues [97]. (B) Glial cells mainly produce N-abridged Aβ peptides (up to 60%) such as Aβ2/3-X or Aβ4/5-x, while neurons produce predominantly Aβ peptides starting at position 1 (80%) [98].
Cells 11 03421 g003
Figure 4. Cellular responses of microglia after detection of Aβ. Microglia are capable of perceiving soluble Aβ peptides and insoluble aggregates via multiple cell-surface receptors, which then induce varied cellular responses such induction of cell migration, cytokine and chemokine release, secretion of proteases, and generation of oxidative stress [3,23]. These processes are in part regulated by the intracellular NLRP3 inflammasome. The release of inflammasome components (ASC specks) may lead to seeding of Aβ and may thus contribute to plaque formation [248].
Figure 4. Cellular responses of microglia after detection of Aβ. Microglia are capable of perceiving soluble Aβ peptides and insoluble aggregates via multiple cell-surface receptors, which then induce varied cellular responses such induction of cell migration, cytokine and chemokine release, secretion of proteases, and generation of oxidative stress [3,23]. These processes are in part regulated by the intracellular NLRP3 inflammasome. The release of inflammasome components (ASC specks) may lead to seeding of Aβ and may thus contribute to plaque formation [248].
Cells 11 03421 g004
Figure 5. Aβ-mediated activation of microglia. Direct detection of Aβ (e.g., by TREM2, TLRs, FPRs, RAGE) or indirect stimulation through Aβ-mediated processes (e.g., TRPM2 activation by ROS, detection of autoantibodies against Aβ by FcRs) leads microglia to adopt their reactive phenotype in which they utilize intracellular signaling pathways and metabolic processes to initiate their pro-inflammatory response.
Figure 5. Aβ-mediated activation of microglia. Direct detection of Aβ (e.g., by TREM2, TLRs, FPRs, RAGE) or indirect stimulation through Aβ-mediated processes (e.g., TRPM2 activation by ROS, detection of autoantibodies against Aβ by FcRs) leads microglia to adopt their reactive phenotype in which they utilize intracellular signaling pathways and metabolic processes to initiate their pro-inflammatory response.
Cells 11 03421 g005
Table 2. Microglial Aβ-binding/binding-modulating receptors and identified AD-associated polymorphisms and mutations. Search for polymorphisms and mutations was conducted using PubMed text search or UniProt https://www.uniprot.org/uniprotkb/.../entry#disease_variants, accessed from 20 June to 4 August 2022. * In the case that no association with AD has been described yet, known polymorphisms or mutations affecting receptor functions are listed.
Table 2. Microglial Aβ-binding/binding-modulating receptors and identified AD-associated polymorphisms and mutations. Search for polymorphisms and mutations was conducted using PubMed text search or UniProt https://www.uniprot.org/uniprotkb/.../entry#disease_variants, accessed from 20 June to 4 August 2022. * In the case that no association with AD has been described yet, known polymorphisms or mutations affecting receptor functions are listed.
ReceptorAlternative Receptor NamesBinding to Aβ Demonstrated inRef.FunctionPolymorphisms/ Mutations Associated with ADRef.
Amylin receptorcalcitonin receptor and receptor activity modifying protein 3human fetal microglial (HFM) cultures and BV-2, oligomeric soluble Aβ1-42[310]Increase IL-6 secretionNo association with AD described
Cys40Trp/Phe100Ser/Leu147Pro variant associated with reduced Amylin potency *
[311]
CD33SIGLEC3No direct binding-Blockade of TREM2rs3865444C[312,313,314,315]
CD36Platelet glycoprotein 4
Fatty acid translocase (FAT)
Glycoprotein IIIb (GPIIIB)
PAS IV
Primary murine microglia, fibrillary Aβ 1-42[235]Increase ROS production and Aβ phagocytosisrs7755 rs3211956
rs3211892
[316]
[317]
CMKLR1 (Chemokine-like receptor 1)Chemerin-like receptor 1 1
G-protein-coupled receptor ChemR23
G-protein-coupled receptor DEZ
stably transfected rat basophilic leukemia (RBL) cells, primary microglia, N9 cells, Aβ-1-42[318]Activation of G proteins and β-arrestin pathwaysNo association with AD described-
FPR1/2 (fMet-Leu-Phe receptors 1/2)N-formyl peptide receptor (FPR)
N-formylpeptide chemoattractant receptor
Rat primary astrocytes and microglia,[319]Intracellular calcium mobilization, cell migration and superoxide anion releaseNo association with AD described-
human Aβ1–42[113]
transfected HEK293 cells and glial U87 cells, Aβ42 and N-truncated forms
MAC1 (Macrophage antigen complex 1)integrin CD11b/CD18 receptor
CR3
Primary microglia enriched culture, mice, Aβ-42[320]Adhesive interactions mediation of the uptake of complement-coated particles and pathogensNo association with AD described
[321]
Various polymorphisms/mutations in CD18 causing leukocyte adhesion deficiency, e.g.,
rs552407409 *
MARCO (Macrophage receptor with collagenous structure)SCARA2Microglia from neonatal rats, fibrillary and non-fibrillary Aβ-42[319,322] Inflammatory responseNo association with AD described-
NMDA-R (NMDA receptor) glial cells in rat cerebellar granule cell cultures, small Aβ1-42 oligomers[323]Decrease in plasma membrane potential-421C/A in sporadic AD, North Han Chinese populations[324]
[325]
3723 G/A (rs3739722), Taiwanese population[326]
rs1806201 T, Southern Italy population[327]
rs10845840, US populations[328]
C2664T, Chinese Han populationnot found in [329]
NucleolinProtein C23EOC2 cells, transfected HEK cells, monomeric and fibrillary Aβ1-42[330]Chromatin decondensation, pre-rRNA transcription, and ribosome assemblyNo association with AD described-
RAGE (Receptor for advanced glycosylation end products)-Human primary microglia, soluble Aβ and plaques[331]instigates pro-inflammatory mediatorsG82S[332,333,334]
cAI (Scavenger receptor type AI)Macrophage scavenger receptor types I and II
Macrophage acetylated LDL receptor I and II
CD204
human fetal microglia, microglia from newborn mice, fibrillary aβ[233]Phagocytosis of soluble and fibrillar AβNo association with AD described-
SRBI (Scavenger receptor class B member 1)CD36 and LIMPII analogous 1 (CLA-1)
CD36 antigen-like 1
Collagen type I receptor, thrombospondin receptor-like 1
human fetal microglia, microglia from newborn mice, fibrillary aβ[233]Decreases amyloid fibrillar and plaque formationGene is included in a region on chromosome 12 with linkage to AD [335], polymorphisms were not found associated[336]
rs387906791, rs74830677, impact on cholesterol metabolism *[337]
[338]
SRCL (scavenger receptor with C-type lectin)Collectin-12
Collectin placenta protein 1 (CL-P1; hCL-P1)
Nurse cell scavenger receptor 2
Scavenger receptor class A member 4
CHO-K1 cells, fibrillary A-β[339]Aβ and Gram-positive, Gram-negative bacteria and yeast phagocytosisNo association with AD described-
TLR2/
TLR4/
CD14
(Toll-like receptor 2/4/6/
CD14)
CD282/CD284/ Monocyte differentiation antigen CD14CD14: soluble murine, fibrillary human Aβ-42[340]Instigates inflammatory response and Aβ phagocytosisP249S (TLR6)
260C/T (CD14)
[341]
[342]
CD14/TLR2/4: BV2 microglia, fibrillary Aβ[343]
TREM1 (Triggering receptor expressed on myeloid cells 1)CD354mouse primary microglia, monomeric Aβ1-42[65]Amplifying inflammatory responsesrs6910730G[344]
TREM2 (Triggering receptor expressed on myeloid cells 2) immobilized TREM2-FC, oligomeric and monomeric Aβ1-42[307]
[345]
Regulates microglial activity, chemotaxis, and process outgrowthR47H
R62H
[346,347,348,349]
[350,351] [352]
[353]
H157Y
D87N
α6β1 integrinCD49 antigen-like family member F
VLA-6 CD49f/ Fibronectin receptor subunit β
Glycoprotein IIa (GPIIA)
VLA-4 subunit β
CD29
BV-2 cells, fibrillary Aβ25-35 and Aβ1-42[354]increase ROS production and Aβ phagocytosisNo association with AD described
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Busch, L.; Eggert, S.; Endres, K.; Bufe, B. The Hidden Role of Non-Canonical Amyloid β Isoforms in Alzheimer’s Disease. Cells 2022, 11, 3421. https://doi.org/10.3390/cells11213421

AMA Style

Busch L, Eggert S, Endres K, Bufe B. The Hidden Role of Non-Canonical Amyloid β Isoforms in Alzheimer’s Disease. Cells. 2022; 11(21):3421. https://doi.org/10.3390/cells11213421

Chicago/Turabian Style

Busch, Lukas, Simone Eggert, Kristina Endres, and Bernd Bufe. 2022. "The Hidden Role of Non-Canonical Amyloid β Isoforms in Alzheimer’s Disease" Cells 11, no. 21: 3421. https://doi.org/10.3390/cells11213421

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop