Next Article in Journal
Relationship between Tumor Infiltrating Immune Cells and Tumor Metastasis and Its Prognostic Value in Cancer
Next Article in Special Issue
Long-Term Memory Formation in Drosophila Depends on the 3′UTR of CPEB Gene orb2
Previous Article in Journal
Myostatin Mutation Enhances Bovine Myogenic Differentiation through PI3K/AKT/mTOR Signalling via Removing DNA Methylation of RACK1
Previous Article in Special Issue
Maternal Hyperhomocysteinemia Produces Memory Deficits Associated with Impairment of Long-Term Synaptic Plasticity in Young Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Roles of Extracellular Vesicles in Alzheimer’s Disease: Focus on Synaptic Dysfunction and Vesicle–Neuron Interaction

1
CNR-Institute of Neuroscience, 20854 Vedano al Lambro, Italy
2
Bio@SNS Laboratory, Scuola Normale Superiore, 56124 Pisa, Italy
3
CNR-Institute of Neuroscience, 56124 Pisa, Italy
*
Authors to whom correspondence should be addressed.
Cells 2023, 12(1), 63; https://doi.org/10.3390/cells12010063
Submission received: 29 November 2022 / Revised: 16 December 2022 / Accepted: 20 December 2022 / Published: 23 December 2022
(This article belongs to the Special Issue New Advances in Synaptic Dysfunctions and Plasticity)

Abstract

:
Alzheimer’s disease (AD) is considered by many to be a synaptic failure. Synaptic function is in fact deeply affected in the very early disease phases and recognized as the main cause of AD-related cognitive impairment. While the reciprocal involvement of amyloid beta (Aβ) and tau peptides in these processes is under intense investigation, the crucial role of extracellular vesicles (EVs) released by different brain cells as vehicles for these molecules and as mediators of early synaptic alterations is gaining more and more ground in the field. In this review, we will summarize the current literature on the contribution of EVs derived from distinct brain cells to neuronal alterations and build a working model for EV-mediated propagation of synaptic dysfunction in early AD. A deeper understanding of EV–neuron interaction will provide useful targets for the development of novel therapeutic approaches aimed at hampering AD progression.

1. Introduction

Alzheimer’s disease is the most common cause of dementia, affecting more than 50 million people worldwide. It is a progressive degenerative encephalopathy causing deterioration in memory, thinking, behaviour, and ability to perform daily activities, and it leads to death. Neuropathological hallmarks of the disease are amyloid beta (Aβ) and tau protein aggregate formation, as well as loss of synapses and neurons [1]. These processes are accompanied by an abnormal response in microglia, the innate immunity cells resident in the brain, which are involved in disease progression [2]. Activated astrocytes, oligodendrocytes, and components of the neurovascular unit also play a role [3]. The neuropathological hallmarks of AD are evident at advanced stages of the disease, first in specific brain regions and then in progressively larger brain areas [4,5,6,7]. By contrast, extensive literature indicates that the first mechanism affected in the disease is functional alteration of the synapses [8,9], which underlies the appearance of clinical symptoms many years later.
Medicine is currently unable to cure AD or to interrupt its progression. Available drugs only treat AD symptoms, temporarily helping memory and thinking problems [10], whereas effective treatments would target the very early stages of the disease, close to synaptic dysfunction onset. For this reason, understanding how synaptic alteration starts and propagates through the brain represents a crucial issue in AD research.
AD pathology progression has been associated with the spreading of Aβ and tau proteins from neuron to neuron throughout the affected brain in animal models [11,12,13,14,15,16]. Importantly, soluble oligomeric forms of these misfolded proteins, before forming extracellular aggregates, have been identified as a cause of synapse dysfunction and neurotoxicity and involved in glia activation. In this context, recent evidence has implicated brain cell-derived extracellular vesicles (EVs) carrying Aβ and tau forms in the onset and propagation of synaptic alterations in AD [15,17,18].
EVs are membrane structures capable of shuttling all sorts of active molecules (proteins, lipids, and nucleic acids) from a donor cell to specific target cells, thus being involved in cell-to-cell communication. EVs can directly bud from the plasma membrane (microvesicles) or be generated in the endocytic compartment (exosomes). Due to technical limitations in isolating and distinguishing EVs depending on their mechanism of biogenesis, the currently recognized nomenclature identifies EVs according to their dimensions, distinguishing medium-large/large EVs (>200 nm) and small EVs (<200 nm) [19].
A large body of evidence indicates that EVs released by neurons and glial cells contain Aβ and tau proteins in transgenic AD mice as well as in culture AD models [15,17,20,21,22,23,24,25,26,27]. The presence of misfolded proteins among EV cargo has been validated in EVs extracted from both body fluids (cerebrospinal fluid (CSF) and plasma) and brain tissue of AD patients [18,20,21,23,28,29,30,31,32]. Encapsulation into EVs protects Aβ and tau from degradation, enhancing their pathogenic action and promoting their diffusion through the brain [15,17,21,23,33].
In this review, we will summarize the current knowledge regarding the role of EVs in the rise and propagation of synaptic alterations in AD, especially as carriers of Aβ and tau proteins, and analyse the possible mechanisms of interaction of EVs with neurons. Findings will be presented based on the cellular source of EVs, to enlighten possible cell-specific EV functions.

2. Synaptic Dysfunction in AD

Synaptic failure is a major determinant of AD [9]. In fact, synapse loss better correlates with cognitive impairment in the disease, rather than with the number of plaques and fibrillary tangles, the grade of neuronal loss, or the extent of gliosis [34]. Furthermore, the synapse is where Aβ peptides are produced and where their effects are targeted [35]. Numerous lines of evidence locate alterations in synaptic function and synapse degeneration in the earliest phases of AD, before the accumulation of misfolded protein aggregates and before neuronal loss [9,34,36].
Aβ oligomers appear to be pivotal players in AD pathogenesis and are particularly active on the synapse [1], being able to alter calcium homeostasis and reduce excitatory synapses’ strength and plasticity. Aβ oligomers impair long-term potentiation (LTP) and long-term depression (LTD), long-term forms of synaptic plasticity thought to underlie learning and memory, in a concentration-dependent manner, leading to cognitive deficits [37]. In particular, high levels of Aβ oligomers affect excitatory synaptic transmission by decreasing the number of surface AMPA and NMDA receptors, dismantling dendritic spines. Compensatory inhibitory responses in circuits involved in learning and memory are also part of AD early synaptic alterations [38]. Changes in synaptic function may cause network instability and lead to synchronous (epileptiform) activity [37]. These effects are not restricted to animal models, as AD patients display aberrant increases in neuronal activity during memory encoding in hypometabolic regions, and early-onset familiar cases show epileptic brain activity.
Pathogenic tau can also be detrimental for the synapse in different ways: it reduces the mobility and release of synaptic vesicles, decreases the number of glutamatergic receptors, affects dendritic spine maturation, disrupts mitochondrial transport and function in the synapses, and promotes the phagocytosis of synapses by microglia [39]. Similarly to Aβ species, tau oligomeric forms are more toxic to the synapse than tau fibrils, impairing LTP and dendritic spine density and maturation, and causing memory deficits [40,41,42]. The regulation of tau by specific phosphorylations recently emerged as crucial for its role at the synapse [43].
Interestingly, several studies support the idea that small oligomeric forms of Aβ and tau may act synergistically in causing synaptic deficits [44,45,46,47,48,49].
Finally, abnormal complement-mediated synaptic pruning seems to play a role in early synaptic loss in AD: complement factors have been found to be necessary for the expression of Aβ effects on the synapse and their depletion to rescue cognitive deficits in APP/PS1 AD mice [50,51,52,53]. Accordingly, it has been proposed that Aβ oligomers enhance the expression of C3 in microglia and astrocytes, driving synapse tagging for elimination [53].

3. EVs and Synaptic Dysfunction in AD

In the brain, EVs are secreted by all cell types, including glial cells and neurons. They can affect the synapse and propagate synaptic alterations among connected cells in a way that poses them as attractive therapeutic targets. The effects of EVs released by different cell types on neuronal function in the context of AD are summarized in Figure 1.

3.1. EVs Released by Microglia

Microglia, the innate immunity cells resident in the brain, are essential regulators of synaptic function and neuronal network formation [2]. They react to the smallest stimulus, being able to assume a various and complex range of activation states [69]. When brain homeostasis is endangered, microglia orchestrate a weighted response to re-establish the status quo [70]. Under sustained brain alterations, as in the case of AD, microglia undergo a neurodegenerative/disease-associated (MGnd/DAM) phenotypic change [71,72,73] and become determinants of disease pathogenesis [74]. Accumulating evidence suggests that DAM might play a positive and protective role in early disease pathology while in late AD stages, DAM might become dysregulated and accelerate the disease [75,76]. The central role of microglia in AD [77] and related synaptic dysfunction [15,50,78,79] has long been known, and the fact that many AD risk genes pertain to microglia and their functions strengthens this concept [77,80,81,82,83].
Studies from our group and others indicated that EVs released by microglia can influence synapse formation, inducing new spines at sites of contact with neurons [17], emulating what happens at microglia–synapse contact sites [84,85]. In addition, cultured neurons exposed to large EVs derived from primary rat microglia show an increase in miniature excitatory postsynaptic current (mEPSC) frequency in a dose-dependent manner, without changes in their amplitude [86]. Analysis of paired-pulse recordings showed that EVs mostly act at the presynaptic site, increasing neurotransmitter release probability [86] and the availability of synaptic vesicles for release [87]. The effect of microglial EVs was confirmed in vivo in cingulate cortex slices from the mouse brain [88], as well as in the rat visual cortex, where injection of large EVs caused an acute increase in the amplitude of field potentials evoked by visual stimuli [86]. Furthermore, a subsequent study showed that microglial large EVs are enriched in endocannabinoids, which are capable of inducing a decrease in miniature inhibitory post-synaptic currents (mIPSCs) targeting CB1 receptors on GABA-ergic cells [89]. More importantly, when microglia are exposed to an inflammatory stimulus, they become detrimental for synaptic function by releasing EVs, which are enriched in a set of miRNAs that regulate the expression of key synaptic proteins. In particular, we demonstrated that large EVs from microglia activated by a cocktail of pro-inflammatory cytokines transfer miR-146a-5p to neurons, leading to the suppression of synaptotagmin 1, a pre-synaptic protein, and neuroligin 1, a postsynaptic adhesion protein that maintains synaptic stability and plays a key role in dendritic spine formation, with detrimental effects on synaptic strength and dendritic spine remodelling [90]. Similar effects on dendritic spines are mediated by small EVs released by primary mouse microglia inflamed after saturated fatty acid palmitate exposure, a model of a high-fat diet [91]. These findings link inflammatory microglia and enhanced EV production to loss of excitatory synapses.
This link was recently confirmed in models of AD, where the implication of large and small microglial EVs in synaptic dysfunction has also been demonstrated.
In a seminal paper, Asai and colleagues used a model of rapid tau propagation from the entorhinal cortex–EC to the dentate gyrus of the hippocampus–DG, together with in vitro systems, to demonstrate that tau propagates between these two regions, causing reduced excitability in DG cells as well as cytopathic changes [15]. Interestingly, tau spreading was limited by both microglia depletion and EV synthesis inhibition [15], while the working and contextual memory deficits were rescued in the P301S tau transgenic mouse model by treatment with a P2 × 7 receptor antagonist, which blocks EV release from microglia [54].
In line with these findings, microglial immune receptor Trem2 deletion in mice (Trem2 KO), a condition known to aggravate tau pathology, enhances tau spreading from the EC to the hippocampus through small EVs, which coincides with impaired synaptic function and memory behaviour [56]. TREM2 is in fact a risk gene for AD and an important regulator of microglia response to pathological changes. R47H heterozygous mutation of TREM2 is linked to late onset AD, and small EVs released by microglia-like cells differentiated from iPSCs in patients carrying this variant (R47Hhet EVs) have been characterized. These EVs contain more inflammatory and DAM-associated proteins than common variant EVs (Cv EVs) [57]; they lose the ability to promote neurite outgrowth and neuronal metabolism; and lose their protective functions against AD-related insults to neurons [58].
Microglia large EVs have been shown to promote the solubilisation of Aβ aggregates, thus shifting equilibrium from an almost inert insoluble form of the peptide toward soluble and neurotoxic species [21,22,92] (Figure 2). In addition, when exposed to Aβ 1–42, microglia release large EVs already carrying neurotoxic Aβ species on their surface and in their lumen [17,21,22]. Once injected into the mouse EC, these Aβ-loaded EVs (Aβ–EVs) reduce synaptic transmission and consequently inhibit LTP. Interestingly, these effects are first detected in the vicinity of the injection site, but synaptic dysfunction then propagates from the EC to the hippocampus [17]. The spreading of synaptic dysfunction was ascribed to the ability of Aβ–EVs to move on the neuron surface, along axonal projections connecting the EC to the DG (Figure 2). Indeed, when Aβ–EV motility was inhibited, no propagation of LTP deficit along the entorhinal–hippocampal circuit occurred [17]. Although it has been reported that tau can be released inside microglia large EVs [93], no data are currently available on the role of such EVs in tau spreading.
Supporting an interplay between tau and Aβ in AD pathogenesis, the release of small tau-carrying EVs is higher from microglia surrounding Aβ plaques than phagocyte hyper-phosphorylated plaque-associated tau, as well as from apoptotic neurons and synapses [55]. Notably, microglia phagocyte AD misfolded proteins and apoptotic structures aiming at their clearance [94], and exploited EV release at least in part as a disposal mechanism [17,21], as other cells do [25,95,96,97].
In accordance with all these studies, large EV production from myeloid cells (microglia/macrophages) is very high in AD patients and correlates with white matter lesions and hippocampal atrophy in prodromal AD, the preeminent expression of neuronal damage in the human brain [98]. In addition, the neurodegenerative microglia signature is enriched in brain-derived small EVs from CAST.APP/PS1 AD mice [30].
Figure 2. Mechanisms of EV-mediated propagation of synaptic dysfunction. Scheme of the onset and propagation of synaptic dysfunction in vulnerable brain regions in the early phases of AD. The model is described in the text section “A model for EV-mediated propagation of synaptic dysfunction”. Parts of the figures were drawn by using pictures from Servier Medical Art, licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/; access date 20 November 2022). Part of the figure is modified from [99], distributed under the terms of the Creative Commons CC-BY license (https://creativecommons.org/licenses/by/4.0/; access date 20 November 2022).
Figure 2. Mechanisms of EV-mediated propagation of synaptic dysfunction. Scheme of the onset and propagation of synaptic dysfunction in vulnerable brain regions in the early phases of AD. The model is described in the text section “A model for EV-mediated propagation of synaptic dysfunction”. Parts of the figures were drawn by using pictures from Servier Medical Art, licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/; access date 20 November 2022). Part of the figure is modified from [99], distributed under the terms of the Creative Commons CC-BY license (https://creativecommons.org/licenses/by/4.0/; access date 20 November 2022).
Cells 12 00063 g002

3.2. EVs Released by Astrocytes

Astrocytes play important roles in neuronal support, maintaining brain homeostasis of ions and neurotransmitters. They represent a fundamental component of the synapse, being part of the so-called “tripartite synapse” together with the pre-synaptic terminal and the post-synaptic compartment [100,101]. Astrocytes are involved in synapse formation, can regulate synaptic transmission, and can also eliminate synapses. Accordingly, similarly to microglial EVs, large EVs released by astrocytes promote excitatory synaptic transmission [86] and move extracellularly, inducing spine formation at sites of stable contact [102], while small EVs carry the neuroprotectant neuroglobin [103], promote neurite outgrowth and neuron survival, and also stimulate synaptic transmission and formation [104,105]. Nevertheless, upon interleukin 1β exposure of donor rat or human primary astrocytes, released small EVs undergo neuronal uptake more frequently than EVs from control cells and are able to inhibit neurite outgrowth, neuronal branching, and firing [104,106].
Like microglia, astrocytes are central players in AD pathology [3] and show early changes in the disease [77,107]. Those close to dystrophic neurites or Aβ plaques alter their morphology, becoming hypertrophic or atrophic [108], as well as their gene and protein expression, displaying a heterogeneous range of activation states [109,110]. In tauopathies, mouse model astrocytes display early functional deficits and lose their neuro-supportive function [111]. In addition, tau accumulation in astrocytes of the DG of the hippocampus, a phenomenon also found in the brain of AD-affected individuals, has been found to cause neuronal dysfunction and memory deficits in mice [112].
Astrocytes are very efficient in engulfing dead cells, synapses, and protein aggregates (e.g., of Aβ) [113,114,115,116,117,118,119], and astrocytes with high Aβ load are frequently found in the AD-affected brain [120]. However, as opposed to microglia, astrocytes are extremely inefficient at degrading phagocytosed material [121], including Aβ 1–42 protofibrils [122]. Aβ accumulation in astrocytes over avery long time further affects endosomal and lysosomal function and induces the release of EVs carrying Aβ (in its N-terminal truncated form) and ApoE to favour elimination of undegraded materials [66,122,123,124]. Furthermore, the Aβ 1–42 proxy Aβ 25–35 induces phosphorilated-tau overproduction in human astrocytes in culture and increases its release within small EVs [24]. EVs carrying Aβ/phosphorylated tau are neurotoxic, causing synaptic loss, axonal swelling, vacuolization of neuronal cell bodies, severe mitochondrial impairment, cholesterol deposits in lysosomal compartments, and apoptosis [66].
The first evidence of the involvement of EVs released by astrocytes in AD progression came from the finding that, in response to Aβ, astrocytes release small EVs containing prostate apoptosis response 4 (PAR4) and ceramide, which induce apoptosis in other astrocytes upon internalization, likely contributing to neurodegeneration [67]. Interestingly, vesicular ceramide was later found to be responsible for astrocyte small EVs’ ability to aggregate Aβ peptides [125] (Figure 2). Subsequently, phosphorylated tau and proteins of the Aβ 1–42 peptide-generating system were found in astrocyte-derived small EVs extracted from the plasma of AD patients [28] as well as various complement proteins that are central players in synaptic pruning [27,126]. In line with this evidence, when isolated from AD patients, astrocyte EVs were more efficient than neuron EVs in inducing complement-mediated neurotoxicity and in reducing neurite density and decreasing cell viability in either cultured neurons or human iPSC-derived neuron-like cells [68].
Additional proofs of the implication of astrocyte EVs in AD progression came from: (i) the enrichment in astrocyte-derived molecules in AD EVs compared to EVs from mild cognitive impairment (MCI) patients [18,127], and (ii) the most significant association of a protein module enriched in astrocyte-specific EV markers with AD pathology and cognitive impairment compared to the proteome of other brain cell-derived AD EVs [128].
Despite extensive evidence suggesting important roles for EVs released by astrocytes in AD synaptopathy evolution, further studies will be necessary to gain a clearer understanding of their early action on the synapse.

3.3. EVs Released by Neurons

During development, neural stem cells can secrete EVs capable of affecting the proliferation and differentiation of neighbouring cells through the propagation of specific miRNAs able to reprogram multiple cellular mechanisms in recipient cells [129,130]. In the mature nervous system, neurons maintain their ability to produce EVs and use these vesicles to communicate with other cells and to regulate several phenomena such as homeostasis, immune response, and synaptic plasticity [131,132]. Although neuronal-derived EVs have been shown to interact with glial cells and affect microglia phagocytic activity [133] and the expression of the glutamate transporter GLT1 in astrocytes [134], in vitro studies suggested that EVs secreted by cortical neurons preferentially bind to other neurons [135], allowing neuron-to-neuron diffusion of specific cargoes. In addition, EV release has been shown to be strongly modulated by synaptic activity [136,137,138].
Neuron-derived EVs could be differentiated from the ones produced by other cell types by the expression of specific markers, such as the L1 cell adhesion molecule (L1CAM), the GluR2/3 subunits of the glutamate receptors, and the GPI-anchored prion protein [136,139]. However, the prion protein was also later identified in astrocyte-derived EVs [102,140].
Given their ability to move from cell to cell, neuronal EVs have been hypothesized to be able to spread along a neural network of connections in a trans-synaptic manner and contribute to the propagation of misfolded proteins in neurodegenerative diseases such as AD.
Indeed, the amyloid precursor protein (APP) and its metabolites, including the Aβ peptide, have been shown to be secreted within neuron-derived small EVs [141,142,143,144,145]. In addition, Sardar Sinha and colleagues [26] demonstrated that the impairment of the formation/secretion of small EVs can suppress the diffusion of Aβ oligomers to other neurons.
Interestingly, as opposed to microglial large EVs, neuronal and neuroblastoma cell line small EVs seem to promote amyloidogenesis of soluble Aβ through the binding of the amyloid peptide to the glycosphingolipid glycans [61,62,92] and to the cellular prion protein (PrPc) [146,147] present on their surface. Neuronal small EVs contain higher levels of glycosphingolipid glycans in their membrane compared to small EVs secreted by other cell types, and this significantly increases Aβ affinity for neuronal-derived EVs [63,148]. The interaction between the Aβ peptide and neuronal EVs can lead to accelerated Aβ fibril formation and, therefore, drive conformational changes in the Aβ to form nontoxic amyloid fibrils [61]. Indeed, small EV markers such as Alix have been observed to be concentrated in senile Aβ plaque in AD brains [141]. Furthermore, PrPC on small EVs negatively regulates Aβ 1–42 uptake by neuronal cells [147] while, on the other hand, neuron-derived small EVs can be efficiently internalized by microglia and promote Aβ degradation, suggesting an overall protective effect of neuron-derived EVs against AD pathology [61,149], as opposed to microglia large EVs [21]. This suggests that neuronal small EVs and microglia large EVs may play very distinct roles in neurodegeneration. In agreement with this hypothesis, protective effects against Aβ-induced pathology and synaptic transmission have been observed following chronic administration of small EVs derived from neuroblastoma cells or primary neurons in the hippocampus [62,63], and a significant rescue of Aβ-induced LTP impairment has been observed after the intracerebroventricular infusion of small EVs in rats [60], strengthening the link between neuronal-derived EVs and neuroprotection. In contrast, EVs released from cultured human neurons and cell lines harbouring familial AD presenilin 1 mutations show neurotoxicity towards cultured wild type neurons in terms of intracellular calcium regulation, mitochondrial functions, and sensibility to excitotoxicity [25].
By means of immunoassays specifically designed to detect the full-length tau protein, considered to be the aggregation-competent form, Guix and colleagues [150] revealed that small EVs secreted by human iPSC-derived neurons or present in human biofluids are highly enriched in full-length tau compared to the extracellular solution, indicating that neuronal EVs carry aggregation-competent tau proteins. In addition, neuronal small EVs could mediate the trans-synaptic propagation of tau protein regardless of its phosphorylation state, in an activity-dependent manner [59], but unfortunately their neurophysiological correlates in vivo have not been explored yet. Interestingly, in analysing neuronal EVs isolated from the plasma of AD and frontotemporal dementia patients, a correlation between the vesicular levels of some synaptic proteins and patients’ cognitive status have been defined, mirroring the decrease in synaptic proteins and the synaptic dysfunction in the affected brain [151,152].

3.4. Mixed EV Populations Isolated from Body Fluids or Brain Tissue

In an increasing number of papers, small EVs isolated from the interstitial space of brain tissue or body fluids (mainly CSF and plasma) have been investigated. These samples represent a real liquid biopsy of the system (animal or human) they are coming from and a window on the microenvironment of specific tissues/compartments in these organisms in a particular situation (e.g., stage of pathology). For this reason, these specimens have been particularly useful for the study of biomarkers for the diagnosis and prognosis of different diseases, including AD and related cognitive defects [27,28,29,31,95,153,154,155,156]. On the other hand, EVs from brain tissue and body fluids are mixed populations of EVs of different cell origins, and the extraction of cell-type-specific EVs is possible only after an additional step of immunoisolation (e.g., in [32]).
As mentioned above (EVs released from astrocytes), small EVs isolated from AD brains typically express more glia- than neuron-derived molecules compared to EVs from healthy subjects [29], and those isolated from the brain of AD mouse models indicate that Aβ can be processed and oligomerized in EVs [157]. Studies on small EVs from human plasma and CSF corroborated the prevalent exposure of Aβ peptides on the surface of EVs [25,158]. A fascinating hypothesis is that binding to the EV surface may be the basis for the low CSF levels of Aβ 1–42 that typically correlate with AD [92]. On the other hand, small oligomeric globular tau, together with other isoforms, phosphorylated or not, have been found inside EVs from tauopathy mice models and AD patients [15,18,23,26,29,31,59] and display an elevated tau seeding activity [18,59,65]. Tau particles have been visualized in the inner leaflet of the EV membranes by electron microscopy [15], and the exposure of even a small portion of tau oligomers on the outer membrane leaflet is highly controversial [18].
Joshi and colleagues were the first to report that EVs isolated from the CSF of AD patients affect neuronal calcium homeostasis and are neurotoxic [21]. A subsequent study showed that small EVs from AD CSF samples are internalized by neurons and affect mitochondrial function, making the cells more vulnerable to excitotoxicity [25]. After internalization, small EVs can be degraded into lysosomes or transfer their content to the cytosol. However, two independent studies recently revealed that small EVs from mouse models and AD patients can avoid disassembly and, still intact, can transport Aβ and tau in an anterograde manner along axons and migrate trans-synaptically to a connected neuron in vitro [26,64] (Figure 2). Interestingly, this might occur also in vivo: small EVs isolated from the plasma of healthy mice and injected into the DG of the hAPP-J20 AD mouse model were engulfed by microglia surrounding Aβ plaques. However, a fraction of them not engulfed by microglia propagated through the hippocampus and up to the cortex in 20 days [159]. Nevertheless, the use of lipophilic dyes to visualize EVs in vivo represents a significant weakness of this study, as explained in the next section.
A first clue that small EVs are able not only to spread throughout the brain, but also to induce and propagate neurophysiological dysfunction in the AD brain came from a study from Dr. Ikezu’s laboratory. In this work, Ruan et al. showed significant spreading of abnormally phosphorylated tau in both the contralateral and ipsilateral hippocampus 4.5 months after inoculation of EVs derived from the brain of prodromal AD and AD patients in the outer membrane layer of the DG [18]. Unexpectedly, tau was mainly found in the GAD67+ interneurons and GluR2/3+ mossy cells in the hilus region of the hippocampus. On the other hand, tau oligomers and fibrils isolated from the same subjects and injected in equal amounts caused very limited tau pathology. Importantly, these phenomena were associated with intrinsic synaptic dysfunction of CA1 pyramidal neurons and reduced input from interneurons and were mediated by tau seeding caused by inoculated EVs.

4. A Model for EV-Mediated Propagation of Synaptic Dysfunction

The study of EV-neuron interaction is fundamental in order to clarify the molecular mechanisms underlying the contribution of EVs to neurodegeneration and to find potential therapeutic targets. Studying the interaction of EVs with target cells in vivo is still difficult due to technical limitations (e.g., low visibility of fluorescent EVs in complex mouse tissue and restricted extracellular space in the brain) and high aspecificity risk associated with EV labelling (i.e., lipophilic dyes such as PKH67, DiR/DiD, MemGlow, and mCling can create EV-like micelles or incidentally label non-EV particles) [160].
The research pioneered by Stefan Momma’s laboratory represents an exception, as his team was able to show the transfer of functional mRNA (Cre mRNA) from hematopoietic EVs to neurons in vivo, taking advantage of a Cre-LoxP recombinase system [161]. Furthermore, the expression of fluorescent EV tags, e.g., CD9-GFP or mEmerald/CD9, under cell-specific promoters in transgenic mice or after lentiviral vector injection is rapidly catching on for in vivo imaging of EVs [55,162], which might include future monitoring of Aβ/tau-carrying EVs in the brain.
However, so far, most of the knowledge on glial EV–neuron interaction comes from in vitro imaging studies, some of which exploit microfluidic or optical manipulation approaches.
Combining such imaging studies with in vivo electrophysiological analyses, we could outline a model for EV–neuron interaction and the spreading of pathological EV cargoes in AD (Figure 2). Following this model, Aβ and tau start to accumulate intracellularly in vulnerable brain regions (1) and are then released by neurons and glia in association with EVs, generating Aβ/tau-carrying EVs (2). In addition, extracellular Aβ and tau bind to lipids at the surface of EVs already present in the pericellular space [61,62,63,92] (3), increasing the pool of Aβ/tau-carrying EVs. Microglial cells can take up Aβ upon endocytosis of neuronal small EVs storing Aβ [61,62] (4) or upon phagocytosis of cytopathic cells or cell debris [55,94] (5), at a later stage.
EVs loaded with Aβ/tau first affect the synapse locally (at the site of EV release or interaction with extracellular Aβ/tau (6)). Then, small EVs are internalized by neurons and travel inside axons, exploiting intracellular trafficking mechanisms [64,163] to trans-synaptically spread their Aβ/tau cargo (as described by [26,59,64] (7)). On the the other hand, larger EVs, too big to be taken up and travel inside axons, move on the neuron surface along axonal projections to reach connected neurons [17,99,102] (8)). Once they reach synaptic sites on dendrites, large EVs can signal locally and/or can be internalized to deliver their cargoes (9). A jump may be necessary for large EVs to transit from the surface of one neuron to a connected one, a phenomenon that occurred in our in vitro studies [102].
This bi-modal mechanism of interaction between small and large EVs and neurons is supported by optical manipulation experiments combined with in vitro fluorescence analysis of EV–neuron contact. Large EVs placed on cultured neurons by optical tweezers and moving on the neuron surface are not internalized by neurons [164] and can be recaptured by the laser trap, which stops their extracellular motion [102]. However, these EVs can undergo partial fusion with the neuron plasma membrane, as indicated by live monitoring of fluorescence intensity of EVs labelled by a dequenching dye, making possible the transfer of EV cargo, such as miRNAs, into neurons [90].
Post-fixation analysis of neurons exposed to fluorescent EVs in bulk provided further information on the fate of small and large glial EVs on neurons: most large EVs (mCLING-labelled) are taken up inside cell somata (94%, [102]) and large dendrites (82%, [102]), but not inside axons (i.e., thin ≤2 μm in diameter neurites in cytoplasmic-RFP-transfected neuronal cultures; 6% astrocyte large EVs, [102]; 3% microglial large EVs, [17]). On the other hand, the internalization of small EVs (lipophilic dye-labelled) of various cell origins into neurons has been largely proven (PKH, [26,33,165]; mCling [140]; DiO, [58]), as well as their transport inside axons for trans-synaptic transfer of their cargo [26,59,64].

Molecular Mechanisms Underlying EV–Neuron Interaction

Insights into the molecular mechanisms underlying EV–neuron interaction come from a few studies. Brenna et al. reported that the cellular prion protein (PrPc), a GPI anchor protein highly enriched in brain EVs [139,140,166], may modulate small EV internalization. Specifically, the presence of PrPc on the surface of astrocyte-derived small EVs limits EV uptake by primary neurons [140]. In agreement with that, we reported that PrPc is enriched on the surface of astrocytic EVs and showed that PrPc interaction with its receptor(s) on neurons favours motion of large EVs at the neuronal surface [102]. In more detail, we found that vesicular PrPc mediates the binding of large EVs to a neuronal receptor (PrPc interacting molecule) that is coupled to a dynamic actin cytoskeleton, thus eliciting passive transport of EVs on neurons [102] (Figure 2 insert A). Only a small fraction of large EVs seem to be able to actively move on the neuron surface, taking advantage of the presence inside their lumen of actin filaments and ATP as their energy source [102,167] (Figure 2 inset B).
A recent study also supported the implication of surface transglutaminase 2 (TG2) in the interaction between astrocytes-derived EVs and neurons [168]. TG2 modulates neuronal calcium homeostasis, synaptic transmission, and cell adhesion. In addition, it interacts with Aβ and other proteins involved in AD, is implicated in Aβ processing, and is over-expressed in AD [169]. Among TG2 interactors involved in cell adhesion, transmembrane proteoglycans, also called syndecans, were shown to mediate adhesion of microglial EVs to neurons (through their heparin sulphate chain [90]). Thus, vesicular TG2 may represent the interactor of neuronal syndecans and contribute to a stable EV–neuron contact.
That glycans have a role in EV cellular uptake is known [170,171]. Galectin is another molecule connected to EV–cell interaction [172], together with integrins, which can bind a plethora of molecules, including intercellular adhesion molecules (ICAMs, [173]) or extracellular matrix (ECM) components on the surface of target cells [174,175,176]. In fact, blocking integrin-mediated interactions partially suppresses astrocyte EV uptake by neurons [106].
Finally, phosphatidylserine (PS), a phospholipid typically exposed on the outer leaflet of the EV membrane [177], also participates in the interaction between glial EVs and neurons. Cloaking PS residues with annexin-V decreases adhesion of astrocytic large EVs to neurons [90] while promoting that of microglial large EVs [17]. Additionally, the treatment largely limits the motion of microglial large EVs in vitro at the neuronal surface and blocks EV-mediated spreading of synaptic dysfunction in vivo, without affecting localized EV action [17].
Proteomic studies in EVs isolated from mouse AD brain cells [30,58,178], human AD specimens [29,32,95,127], and iPSC-derived neural cells [128] confirmed changes in the expression of adhesion molecules (PrPC, integrin-β1, -β2, -αx, annexin A5, A6, A7, neuroligin, Thy-1, CCL2, APP, and ApoE), and also showed alterations in molecules involved in actin cytoskeleton dynamics (glicoprotein M6-A, formin-like protein-1, the puromycin sensitive aminopeptidase interacting with ezrin, and an ERM protein crosslinking actin filaments with plasma membranes), or those interacting with microtubules (e.g., α-synuclein [179]). Canonical pathways of small EV proteins from human AD CSF include various endocytic pathways, virus entry, cell adhesion, and actin cytoskeleton signalling [127]. Interestingly, APP and PrPc, which are able to interact with each other and with many other proteins, including Aβ, tau, the ECM, adhesion molecules, cell cytoskeleton [41,180,181,182,183,184,185], are significantly co-upregulated in brain EVs from AD patients compared to age-matched controls specifically in the preclinical stages of the disease, suggesting their implication in disease onset and/or early development [95]. Notably, APP has been reported as necessary for Aβ and tau oligomers to enter the neurons and induce abnormal synaptic function and memory [41,186].
Therefore, all the adhesion molecules interacting with the cell cytoskeleton that are differentially expressed in AD EVs may influence EV–neuron contact and signalling. In accordance, small EVs isolated from the frontal cortex of AD patients display an increased uptake by neurons (coupled with enhanced tau transfer efficacy) compared to EVs from prodromal AD and control subjects [18]. In contrast, microglial large EVs carrying Aβ are more prone to extracellular motion on axons, and they move faster and preferentially towards the cell periphery compared to EVs not carrying Aβ [17]. However, it remains largely unknown which surface proteins are implicated in the alteration of EV–neuron interactions.
It should be also taken into consideration that, to move extracellularly in the brain parenchyma, EVs have to interact with the ECM. Not surprisingly, several EV surface molecules act on or interact with the ECM [187,188,189,190] (e.g., metalloprotease and others), while ECM components can associate with the EV membrane [191]. The ECM is involved in several brain processes, including synaptic plasticity [192], and is altered in AD, especially regarding the perineuronal nets that form net-like structures around neurons [35], thus potentially affecting EV extracellular motion.

5. Discussion

Loss of synaptic functionality is the earliest event in AD pathology. However, little is known about how synaptic dysfunction starts and propagates along neuronal circuits to cause network abnormalities. The role of EVs of different origins as vehicles for Aβ and tau spreading throughout the brain and across the synapse is widely accepted. So far, only a few papers have examined the ability of EVs to effectively induce and propagate AD-related neurophysiological alterations in the brain [15,17,18]. These studies highlight a major role for both small and large EVs in these processes, placing them at the top of the list of putative targets for novel therapies.
While we still lack a direct proof of EV motion in vivo, in vitro studies provided us with some clues to build a model for EV-mediated propagation of Aβ/tau and related synaptic deficits (Figure 2). In this model, large and small EVs interact with neurons in distinct ways, spreading dysfunction by moving at the axonal surface or inside neuronal processes, respectively.
The challenge is now to decipher the molecular mechanisms underlying EV–neuron interaction and EV motion inside/outside neuronal processes. This will represent a significant step towards the establishment of novel therapeutic strategies to hamper AD progression. Notably, these mechanisms may be shared by other neurodegenerative diseases characterized by the spreading of misfolded proteins along neuronal connections, such as Parkinson’s disease and amyotrophic lateral sclerosis (ALS).
It is worth noting that, while the current knowledge about EV action on the synapse in AD comes from the analysis of EVs of either multiple cell origins or derived from microglia and only marginally astrocytes and neurons, still unexplored is the role of EVs released by oligodendrocytes, which typically exert neuroprotective functions aimed at preserving brain homeostasis [193], and the role of other brain cells, such as vascular and blood–brain–barrier (BBB) epithelium cells, which may play some roles in the evolution of the pathology. Furthermore, the large majority of the studies regarding AD focus on small EVs, although recent evidence reveals that large EVs deserve more attention as feasible pharmacological targets. In addition, Aβ-tau interplay, which notably magnifies neuronal circuit impairment [47], is something that would be worth looking into more deeply in relation to AD EV action.
Relative to the mechanisms of EV-mediated induction of synaptic failure, it should be pointed out that EVs may cross the BBB [161], and systemic inflammation may also modulate synaptic function via blood EVs. This is somewhat suggested by the evidence that intracerebroventricular injection of intestinal small EVs from intestinal ischemia/reperfusion mice causes microglial activation, neuronal loss, synaptic stability decline, and cognitive impairment [194]. Moreover, the exposure of neurons to EVs derived from systemic autoimmune-driven diseases such as rheumatoid arthritis leads to synaptic dysfunction, as evidenced by electrophysiological recordings of mEPSCs and analysis of synaptic markers [195]. Investigating the mechanisms regulating EV entrance in the brain will be instrumental to clarify the role of peripheral EVs in AD synaptic failure.
Finally, small EVs of different origin have been proven as potential therapeutic agents in AD models. Aside from the beneficial action of neural and mesenchymal stem cell-derived small EVs, which mirror the well-known immunomodulatory and therapeutic power of their donor cells [196], neuronal EVs showed protective and curative effects against Aβ-induced pathology and synaptic alterations. In addition, understanding how EVs of oligodendrocyte and vascular cell origin influence synaptic transmission in AD may provide new options for EV-mediated therapeutic applications in the disease.

6. Conclusions

To conclude, recent findings point to a substantial role for large and small EVs of different cell sources in the onset and propagation of synaptic alterations in early AD. Given the lack of effective therapies for this terrible disease, decoding the molecular mechanisms at the basis of the biological action of EV will be crucial in developing novel strategies to limit the progression of neurodegeneration.

Author Contributions

Conceptualization, M.G., F.T., C.V. and N.O.; Writing–Original Draft Preparation, Review and Editing, M.G., F.T., C.V. and N.O. Figures, M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of University and Research-PNRR project n. (T.H.E.) ECS_00000017 to N.O., Alzheimer’s Association Research Fellowship (AARF) 2018-AARF-588984, and the Italian Ministry of Health (RF-2018-12365333) to C.V.

Acknowledgments

Parts of the figures were drawn by using pictures from Servier Medical Art, licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/; access date 20 November 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Selkoe, D.J.; Hardy, J. The amyloid hypothesis of Alzheimer’s disease at 25 years. EMBO Mol. Med. 2016, 8, 595–608. [Google Scholar] [CrossRef] [PubMed]
  2. Salter, M.W.; Stevens, B. Microglia Emerge as Central Players in Brain Disease. Nat. Med. 2017, 23, 1018–1027. [Google Scholar] [CrossRef] [PubMed]
  3. De Strooper, B.; Karran, E. The Cellular Phase of Alzheimer’s Disease. Cell 2016, 164, 603–615. [Google Scholar] [CrossRef] [Green Version]
  4. Braak, H.; Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991, 82, 239–259. [Google Scholar] [CrossRef] [PubMed]
  5. Thal, D.R.; Rub, U.; Orantes, M.; Braak, H. Phases of A beta-deposition in the human brain and its relevance for the development of AD. Neurology 2002, 58, 1791–1800. [Google Scholar] [CrossRef]
  6. Braak, H.; Alafuzoff, I.; Arzberger, T.; Kretzschmar, H.; Del Tredici, K. Staging of Alzheimer disease-associated neurofibrillary pathology using paraffin sections and immunocytochemistry. Acta Neuropathol. 2006, 112, 389–404. [Google Scholar] [CrossRef] [Green Version]
  7. Sepulcre, J.; Masdeu, J.C. Advanced Neuroimaging Methods Towards Characterization of Early Stages of Alzheimer’s Disease. Methods Mol. Biol. 2016, 1303, 509–519. [Google Scholar] [CrossRef]
  8. Masliah, E. Mechanisms of synaptic dysfunction in Alzheimer’s disease. Histol. Histopathol. 1995, 10, 509–519. [Google Scholar] [PubMed]
  9. Selkoe, D.J. Alzheimer’s disease is a synaptic failure. Science 2002, 298, 789–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Godyn, J.; Jonczyk, J.; Panek, D.; Malawska, B. Therapeutic strategies for Alzheimer’s disease in clinical trials. Pharmacol. Rep. 2016, 68, 127–138. [Google Scholar] [CrossRef]
  11. Dujardin, S.; Lecolle, K.; Caillierez, R.; Begard, S.; Zommer, N.; Lachaud, C.; Carrier, S.; Dufour, N.; Auregan, G.; Winderickx, J.; et al. Neuron-to-neuron wild-type Tau protein transfer through a trans-synaptic mechanism: Relevance to sporadic tauopathies. Acta Neuropathol. Commun. 2014, 2, 14. [Google Scholar] [CrossRef] [PubMed]
  12. Harris, J.A.; Devidze, N.; Verret, L.; Ho, K.; Halabisky, B.; Thwin, M.T.; Kim, D.; Hamto, P.; Lo, I.; Yu, G.Q.; et al. Transsynaptic progression of amyloid-beta-induced neuronal dysfunction within the entorhinal-hippocampal network. Neuron 2010, 68, 428–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Liu, L.; Drouet, V.; Wu, J.W.; Witter, M.P.; Small, S.A.; Clelland, C.; Duff, K. Trans-synaptic spread of tau pathology in vivo. PLoS ONE 2012, 7, e31302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Takeda, S.; Wegmann, S.; Cho, H.; DeVos, S.L.; Commins, C.; Roe, A.D.; Nicholls, S.B.; Carlson, G.A.; Pitstick, R.; Nobuhara, C.K.; et al. Neuronal uptake and propagation of a rare phosphorylated high-molecular-weight tau derived from Alzheimer’s disease brain. Nat. Commun. 2015, 6, 8490. [Google Scholar] [CrossRef] [Green Version]
  15. Asai, H.; Ikezu, S.; Tsunoda, S.; Medalla, M.; Luebke, J.; Haydar, T.; Wolozin, B.; Butovsky, O.; Kugler, S.; Ikezu, T. Depletion of microglia and inhibition of exosome synthesis halt tau propagation. Nat. Neurosci. 2015, 18, 1584–1593. [Google Scholar] [CrossRef] [Green Version]
  16. Delpech, J.C.; Pathak, D.; Varghese, M.; Kalavai, S.V.; Hays, E.C.; Hof, P.R.; Johnson, W.E.; Ikezu, S.; Medalla, M.; Luebke, J.I.; et al. Wolframin-1-expressing neurons in the entorhinal cortex propagate tau to CA1 neurons and impair hippocampal memory in mice. Sci. Transl. Med. 2021, 13, eabe8455. [Google Scholar] [CrossRef]
  17. Gabrielli, M.; Prada, I.; Joshi, P.; Falcicchia, C.; D’Arrigo, G.; Rutigliano, G.; Battocchio, E.; Zenatelli, R.; Tozzi, F.; Radeghieri, A.; et al. Microglial large extracellular vesicles propagate early synaptic dysfunction in Alzheimer’s disease. Brain 2022, 145, 2849–2868. [Google Scholar] [CrossRef]
  18. Ruan, Z.; Pathak, D.; Venkatesan Kalavai, S.; Yoshii-Kitahara, A.; Muraoka, S.; Bhatt, N.; Takamatsu-Yukawa, K.; Hu, J.; Wang, Y.; Hersh, S.; et al. Alzheimer’s disease brain-derived extracellular vesicles spread tau pathology in interneurons. Brain 2021, 144, 288–309. [Google Scholar] [CrossRef]
  19. Thery, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef] [Green Version]
  20. Saman, S.; Kim, W.; Raya, M.; Visnick, Y.; Miro, S.; Saman, S.; Jackson, B.; McKee, A.C.; Alvarez, V.E.; Lee, N.C.; et al. Exosome-associated tau is secreted in tauopathy models and is selectively phosphorylated in cerebrospinal fluid in early Alzheimer disease. J. Biol. Chem. 2012, 287, 3842–3849. [Google Scholar] [CrossRef]
  21. 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] [PubMed] [Green Version]
  22. Gouwens, L.K.; Ismail, M.S.; Rogers, V.A.; Zeller, N.T.; Garrad, E.C.; Amtashar, F.S.; Makoni, N.J.; Osborn, D.C.; Nichols, M.R. Abeta42 Protofibrils Interact with and Are Trafficked through Microglial-Derived Microvesicles. ACS Chem. Neurosci. 2018, 9, 1416–1425. [Google Scholar] [CrossRef] [PubMed]
  23. Crotti, A.; Sait, H.R.; McAvoy, K.M.; Estrada, K.; Ergun, A.; Szak, S.; Marsh, G.; Jandreski, L.; Peterson, M.; Reynolds, T.L.; et al. BIN1 favors the spreading of Tau via extracellular vesicles. Sci. Rep. 2019, 9, 9477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Chiarini, A.; Armato, U.; Gardenal, E.; Gui, L.; Dal Pra, I. Amyloid beta-Exposed Human Astrocytes Overproduce Phospho-Tau and Overrelease It within Exosomes, Effects Suppressed by Calcilytic NPS 2143-Further Implications for Alzheimer’s Therapy. Front. Neurosci. 2017, 11, 217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Eitan, E.; Hutchison, E.R.; Marosi, K.; Comotto, J.; Mustapic, M.; Nigam, S.M.; Suire, C.; Maharana, C.; Jicha, G.A.; Liu, D.; et al. Extracellular Vesicle-Associated Abeta Mediates Trans-Neuronal Bioenergetic and Ca2+-Handling Deficits in Alzheimer’s Disease Models. NPJ Aging Mech. Dis. 2016, 2, 16019. [Google Scholar] [CrossRef] [Green Version]
  26. Sardar Sinha, M.; Ansell-Schultz, A.; Civitelli, L.; Hildesjo, C.; Larsson, M.; Lannfelt, L.; Ingelsson, M.; Hallbeck, M. Alzheimer’s disease pathology propagation by exosomes containing toxic amyloid-beta oligomers. Acta Neuropathol. 2018, 136, 41–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Delgado-Peraza, F.; Nogueras-Ortiz, C.J.; Volpert, O.; Liu, D.; Goetzl, E.J.; Mattson, M.P.; Greig, N.H.; Eitan, E.; Kapogiannis, D. Neuronal and Astrocytic Extracellular Vesicle Biomarkers in Blood Reflect Brain Pathology in Mouse Models of Alzheimer’s Disease. Cells 2021, 10, 993. [Google Scholar] [CrossRef] [PubMed]
  28. Goetzl, E.J.; Mustapic, M.; Kapogiannis, D.; Eitan, E.; Lobach, I.V.; Goetzl, L.; Schwartz, J.B.; Miller, B.L. Cargo proteins of plasma astrocyte-derived exosomes in Alzheimer’s disease. FASEB J. 2016, 30, 3853–3859. [Google Scholar] [CrossRef] [Green Version]
  29. Muraoka, S.; DeLeo, A.M.; Sethi, M.K.; Yukawa-Takamatsu, K.; Yang, Z.; Ko, J.; Hogan, J.D.; Ruan, Z.; You, Y.; Wang, Y.; et al. Proteomic and biological profiling of extracellular vesicles from Alzheimer’s disease human brain tissues. Alzheimer’s Dement. 2020, 16, 896–907. [Google Scholar] [CrossRef]
  30. Muraoka, S.; Jedrychowski, M.P.; Iwahara, N.; Abdullah, M.; Onos, K.D.; Keezer, K.J.; Hu, J.; Ikezu, S.; Howell, G.R.; Gygi, S.P.; et al. Enrichment of Neurodegenerative Microglia Signature in Brain-Derived Extracellular Vesicles Isolated from Alzheimer’s Disease Mouse Models. J. Proteome Res. 2021, 20, 1733–1743. [Google Scholar] [CrossRef]
  31. Fiandaca, M.S.; Kapogiannis, D.; Mapstone, M.; Boxer, A.; Eitan, E.; Schwartz, J.B.; Abner, E.L.; Petersen, R.C.; Federoff, H.J.; Miller, B.L.; et al. Identification of preclinical Alzheimer’s disease by a profile of pathogenic proteins in neurally derived blood exosomes: A case-control study. Alzheimer’s Dement. 2015, 11, 600–607.e601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. 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. Pharmacol. 2021, 12, 766082. [Google Scholar] [CrossRef] [PubMed]
  33. Guo, M.; Wang, J.; Zhao, Y.; Feng, Y.; Han, S.; Dong, Q.; Cui, M.; Tieu, K. Microglial exosomes facilitate alpha-synuclein transmission in Parkinson’s disease. Brain 2020, 143, 1476–1497. [Google Scholar] [CrossRef] [PubMed]
  34. Terry, R.D.; Masliah, E.; Salmon, D.P.; Butters, N.; DeTeresa, R.; Hill, R.; Hansen, L.A.; Katzman, R. Physical basis of cognitive alterations in Alzheimer’s disease: Synapse loss is the major correlate of cognitive impairment. Ann. Neurol. 1991, 30, 572–580. [Google Scholar] [CrossRef] [PubMed]
  35. Pelucchi, S.; Gardoni, F.; Di Luca, M.; Marcello, E. Synaptic dysfunction in early phases of Alzheimer’s Disease. Handb. Clin. Neurol. 2022, 184, 417–438. [Google Scholar] [CrossRef]
  36. Masliah, E.; Mallory, M.; Alford, M.; DeTeresa, R.; Hansen, L.A.; McKeel, D.W., Jr.; Morris, J.C. Altered expression of synaptic proteins occurs early during progression of Alzheimer’s disease. Neurology 2001, 56, 127–129. [Google Scholar] [CrossRef] [Green Version]
  37. Palop, J.J.; Mucke, L. Amyloid-beta-induced neuronal dysfunction in Alzheimer’s disease: From synapses toward neural networks. Nat. Neurosci. 2010, 13, 812–818. [Google Scholar] [CrossRef] [Green Version]
  38. Palop, J.J.; Chin, J.; Roberson, E.D.; Wang, J.; Thwin, M.T.; Bien-Ly, N.; Yoo, J.; Ho, K.O.; Yu, G.Q.; Kreitzer, A.; et al. Aberrant excitatory neuronal activity and compensatory remodeling of inhibitory hippocampal circuits in mouse models of Alzheimer’s disease. Neuron 2007, 55, 697–711. [Google Scholar] [CrossRef] [Green Version]
  39. Wu, M.; Zhang, M.; Yin, X.; Chen, K.; Hu, Z.; Zhou, Q.; Cao, X.; Chen, Z.; Liu, D. The role of pathological tau in synaptic dysfunction in Alzheimer’s diseases. Transl. Neurodegener. 2021, 10, 45. [Google Scholar] [CrossRef]
  40. Fa, M.; Puzzo, D.; Piacentini, R.; Staniszewski, A.; Zhang, H.; Baltrons, M.A.; Li Puma, D.D.; Chatterjee, I.; Li, J.; Saeed, F.; et al. Extracellular Tau Oligomers Produce An Immediate Impairment of LTP and Memory. Sci. Rep. 2016, 6, 19393. [Google Scholar] [CrossRef]
  41. Puzzo, D.; Piacentini, R.; Fa, M.; Gulisano, W.; Li Puma, D.D.; Staniszewski, A.; Zhang, H.; Tropea, M.R.; Cocco, S.; Palmeri, A.; et al. LTP and memory impairment caused by extracellular Abeta and Tau oligomers is APP-dependent. eLife 2017, 6, e26991. [Google Scholar] [CrossRef] [PubMed]
  42. Kaniyappan, S.; Chandupatla, R.R.; Mandelkow, E.M.; Mandelkow, E. Extracellular low-n oligomers of tau cause selective synaptotoxicity without affecting cell viability. Alzheimer’s Dement. 2017, 13, 1270–1291. [Google Scholar] [CrossRef] [PubMed]
  43. Regan, P.; Whitcomb, D.J.; Cho, K. Physiological and Pathophysiological Implications of Synaptic Tau. Neuroscientist 2017, 23, 137–151. [Google Scholar] [CrossRef] [Green Version]
  44. Ittner, L.M.; Gotz, J. Amyloid-beta and tau—A toxic pas de deux in Alzheimer’s disease. Nat. Rev. Neurosci. 2011, 12, 65–72. [Google Scholar] [CrossRef] [PubMed]
  45. Roberson, E.D.; Scearce-Levie, K.; Palop, J.J.; Yan, F.; Cheng, I.H.; Wu, T.; Gerstein, H.; Yu, G.Q.; Mucke, L. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer’s disease mouse model. Science 2007, 316, 750–754. [Google Scholar] [CrossRef] [Green Version]
  46. Sperling, R.A.; Mormino, E.C.; Schultz, A.P.; Betensky, R.A.; Papp, K.V.; Amariglio, R.E.; Hanseeuw, B.J.; Buckley, R.; Chhatwal, J.; Hedden, T.; et al. The impact of amyloid-beta and tau on prospective cognitive decline in older individuals. Ann. Neurol. 2019, 85, 181–193. [Google Scholar] [CrossRef] [PubMed]
  47. Busche, M.A.; Hyman, B.T. Synergy between amyloid-beta and tau in Alzheimer’s disease. Nat. Neurosci. 2020, 23, 1183–1193. [Google Scholar] [CrossRef]
  48. Crimins, J.L.; Pooler, A.; Polydoro, M.; Luebke, J.I.; Spires-Jones, T.L. The intersection of amyloid beta and tau in glutamatergic synaptic dysfunction and collapse in Alzheimer’s disease. Ageing Res. Rev. 2013, 12, 757–763. [Google Scholar] [CrossRef] [Green Version]
  49. Spires-Jones, T.L.; Hyman, B.T. The intersection of amyloid beta and tau at synapses in Alzheimer’s disease. Neuron 2014, 82, 756–771. [Google Scholar] [CrossRef] [Green Version]
  50. Hong, S.; Beja-Glasser, V.F.; Nfonoyim, B.M.; Frouin, A.; Li, S.; Ramakrishnan, S.; Merry, K.M.; Shi, Q.; Rosenthal, A.; Barres, B.A.; et al. Complement and microglia mediate early synapse loss in Alzheimer mouse models. Science 2016, 352, 712–716. [Google Scholar] [CrossRef]
  51. Fonseca, M.I.; Zhou, J.; Botto, M.; Tenner, A.J. Absence of C1q leads to less neuropathology in transgenic mouse models of Alzheimer’s disease. J. Neurosci. 2004, 24, 6457–6465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Shi, Q.; Chowdhury, S.; Ma, R.; Le, K.X.; Hong, S.; Caldarone, B.J.; Stevens, B.; Lemere, C.A. Complement C3 deficiency protects against neurodegeneration in aged plaque-rich APP/PS1 mice. Sci. Transl. Med. 2017, 9, eaaf6295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Rajendran, L.; Paolicelli, R.C. Microglia-Mediated Synapse Loss in Alzheimer’s Disease. J. Neurosci. 2018, 38, 2911–2919. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ruan, Z.; Delpech, J.C.; Venkatesan Kalavai, S.; Van Enoo, A.A.; Hu, J.; Ikezu, S.; Ikezu, T. P2RX7 inhibitor suppresses exosome secretion and disease phenotype in P301S tau transgenic mice. Mol. Neurodegener. 2020, 15, 47. [Google Scholar] [CrossRef] [PubMed]
  55. Clayton, K.; Delpech, J.C.; Herron, S.; Iwahara, N.; Ericsson, M.; Saito, T.; Saido, T.C.; Ikezu, S.; Ikezu, T. Plaque associated microglia hyper-secrete extracellular vesicles and accelerate tau propagation in a humanized APP mouse model. Mol. Neurodegener. 2021, 16, 18. [Google Scholar] [CrossRef]
  56. Zhu, B.; Liu, Y.; Hwang, S.; Archuleta, K.; Huang, H.; Campos, A.; Murad, R.; Pina-Crespo, J.; Xu, H.; Huang, T.Y. Trem2 deletion enhances tau dispersion and pathology through microglia exosomes. Mol. Neurodegener. 2022, 17, 58. [Google Scholar] [CrossRef]
  57. Mallach, A.; Gobom, J.; Arber, C.; Piers, T.M.; Hardy, J.; Wray, S.; Zetterberg, H.; Pocock, J. Differential Stimulation of Pluripotent Stem Cell-Derived Human Microglia Leads to Exosomal Proteomic Changes Affecting Neurons. Cells 2021, 10, 2866. [Google Scholar] [CrossRef]
  58. Mallach, A.; Gobom, J.; Zetterberg, H.; Hardy, J.; Piers, T.M.; Wray, S.; Pocock, J.M. The influence of the R47H triggering receptor expressed on myeloid cells 2 variant on microglial exosome profiles. Brain Commun. 2021, 3, fcab009. [Google Scholar] [CrossRef]
  59. Wang, Y.; Balaji, V.; Kaniyappan, S.; Kruger, L.; Irsen, S.; Tepper, K.; Chandupatla, R.; Maetzler, W.; Schneider, A.; Mandelkow, E.; et al. The release and trans-synaptic transmission of Tau via exosomes. Mol. Neurodegener. 2017, 12, 5. [Google Scholar] [CrossRef] [Green Version]
  60. An, K.; Klyubin, I.; Kim, Y.; Jung, J.H.; Mably, A.J.; O’Dowd, S.T.; Lynch, T.; Kanmert, D.; Lemere, C.A.; Finan, G.M.; et al. Exosomes neutralize synaptic-plasticity-disrupting activity of Abeta assemblies in vivo. Mol. Brain 2013, 6, 47. [Google Scholar] [CrossRef]
  61. Yuyama, K.; Sun, H.; Mitsutake, S.; Igarashi, Y. Sphingolipid-modulated exosome secretion promotes clearance of amyloid-beta by microglia. J. Biol. Chem. 2012, 287, 10977–10989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Yuyama, K.; Sun, H.; Sakai, S.; Mitsutake, S.; Okada, M.; Tahara, H.; Furukawa, J.; Fujitani, N.; Shinohara, Y.; Igarashi, Y. Decreased amyloid-beta pathologies by intracerebral loading of glycosphingolipid-enriched exosomes in Alzheimer model mice. J. Biol. Chem. 2014, 289, 24488–24498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Yuyama, K.; Sun, H.; Usuki, S.; Sakai, S.; Hanamatsu, H.; Mioka, T.; Kimura, N.; Okada, M.; Tahara, H.; Furukawa, J.; et al. A potential function for neuronal exosomes: Sequestering intracerebral amyloid-beta peptide. FEBS Lett. 2015, 589, 84–88. [Google Scholar] [CrossRef] [PubMed]
  64. Polanco, J.C.; Li, C.; Durisic, N.; Sullivan, R.; Gotz, J. Exosomes taken up by neurons hijack the endosomal pathway to spread to interconnected neurons. Acta Neuropathol. Commun. 2018, 6, 10. [Google Scholar] [CrossRef] [Green Version]
  65. Leroux, E.; Perbet, R.; Caillierez, R.; Richetin, K.; Lieger, S.; Espourteille, J.; Bouillet, T.; Begard, S.; Danis, C.; Loyens, A.; et al. Extracellular vesicles: Major actors of heterogeneity in tau spreading among human tauopathies. Mol. Ther. 2022, 30, 782–797. [Google Scholar] [CrossRef]
  66. Beretta, C.; Nikitidou, E.; Streubel-Gallasch, L.; Ingelsson, M.; Sehlin, D.; Erlandsson, A. Extracellular vesicles from amyloid-beta exposed cell cultures induce severe dysfunction in cortical neurons. Sci. Rep. 2020, 10, 19656. [Google Scholar] [CrossRef]
  67. Wang, G.; Dinkins, M.; He, Q.; Zhu, G.; Poirier, C.; Campbell, A.; Mayer-Proschel, M.; Bieberich, E. Astrocytes secrete exosomes enriched with proapoptotic ceramide and prostate apoptosis response 4 (PAR-4): Potential mechanism of apoptosis induction in Alzheimer disease (AD). J. Biol. Chem. 2012, 287, 21384–21395. [Google Scholar] [CrossRef] [Green Version]
  68. Nogueras-Ortiz, C.J.; Mahairaki, V.; Delgado-Peraza, F.; Das, D.; Avgerinos, K.; Eren, E.; Hentschel, M.; Goetzl, E.J.; Mattson, M.P.; Kapogiannis, D. Astrocyte- and Neuron-Derived Extracellular Vesicles from Alzheimer’s Disease Patients Effect Complement-Mediated Neurotoxicity. Cells 2020, 9, 1618. [Google Scholar] [CrossRef]
  69. Paolicelli, R.C.; Sierra, A.; Stevens, B.; Tremblay, M.E.; Aguzzi, A.; Ajami, B.; Amit, I.; Audinat, E.; Bechmann, I.; Bennett, M.; et al. Microglia states and nomenclature: A field at its crossroads. Neuron 2022, 110, 3458–3483. [Google Scholar] [CrossRef]
  70. Cserep, C.; Posfai, B.; Lenart, N.; Fekete, R.; Laszlo, Z.I.; Lele, Z.; Orsolits, B.; Molnar, G.; Heindl, S.; Schwarcz, A.D.; et al. Microglia monitor and protect neuronal function through specialized somatic purinergic junctions. Science 2020, 367, 528–537. [Google Scholar] [CrossRef]
  71. 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.e1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Krasemann, S.; Madore, C.; Cialic, R.; Baufeld, C.; Calcagno, N.; El Fatimy, R.; Beckers, L.; O’Loughlin, E.; Xu, Y.; Fanek, Z.; et al. The TREM2-APOE Pathway Drives the Transcriptional Phenotype of Dysfunctional Microglia in Neurodegenerative Diseases. Immunity 2017, 47, 566–581.e569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Mathys, H.; Adaikkan, C.; Gao, F.; Young, J.Z.; Manet, E.; Hemberg, M.; De Jager, P.L.; Ransohoff, R.M.; Regev, A.; Tsai, L.H. Temporal Tracking of Microglia Activation in Neurodegeneration at Single-Cell Resolution. Cell Rep. 2017, 21, 366–380. [Google Scholar] [CrossRef] [Green Version]
  74. 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]
  75. Jay, T.R.; Hirsch, A.M.; Broihier, M.L.; Miller, C.M.; Neilson, L.E.; Ransohoff, R.M.; Lamb, B.T.; Landreth, G.E. Disease Progression-Dependent Effects of TREM2 Deficiency in a Mouse Model of Alzheimer’s Disease. J. Neurosci. 2017, 37, 637–647. [Google Scholar] [CrossRef] [PubMed]
  76. Deczkowska, A.; Weiner, A.; Amit, I. The Physiology, Pathology, and Potential Therapeutic Applications of the TREM2 Signaling Pathway. Cell 2020, 181, 1207–1217. [Google Scholar] [CrossRef] [PubMed]
  77. Heneka, M.T.; Carson, M.J.; El Khoury, J.; Landreth, G.E.; Brosseron, F.; Feinstein, D.L.; Jacobs, A.H.; Wyss-Coray, T.; Vitorica, J.; Ransohoff, R.M.; et al. Neuroinflammation in Alzheimer’s disease. Lancet Neurol. 2015, 14, 388–405. [Google Scholar] [CrossRef] [Green Version]
  78. 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]
  79. Origlia, N.; Bonadonna, C.; Rosellini, A.; Leznik, E.; Arancio, O.; Yan, S.S.; Domenici, L. Microglial receptor for advanced glycation end product-dependent signal pathway drives beta-amyloid-induced synaptic depression and long-term depression impairment in entorhinal cortex. J. Neurosci. 2010, 30, 11414–11425. [Google Scholar] [CrossRef] [Green Version]
  80. Guerreiro, R.; Wojtas, A.; Bras, J.; Carrasquillo, M.; Rogaeva, E.; Majounie, E.; Cruchaga, C.; Sassi, C.; Kauwe, J.S.; Younkin, S.; et al. TREM2 variants in Alzheimer’s disease. N. Engl. J. Med. 2013, 368, 117–127. [Google Scholar] [CrossRef]
  81. Jonsson, T.; Stefansson, H.; Steinberg, S.; Jonsdottir, I.; Jonsson, P.V.; Snaedal, J.; Bjornsson, S.; Huttenlocher, J.; Levey, A.I.; Lah, J.J.; et al. Variant of TREM2 associated with the risk of Alzheimer’s disease. N. Engl. J. Med. 2013, 368, 107–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhang, B.; Gaiteri, C.; Bodea, L.G.; Wang, Z.; McElwee, J.; Podtelezhnikov, A.A.; Zhang, C.; Xie, T.; Tran, L.; Dobrin, R.; et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell 2013, 153, 707–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Lambert, J.C.; Ibrahim-Verbaas, C.A.; Harold, D.; Naj, A.C.; Sims, R.; Bellenguez, C.; DeStafano, A.L.; Bis, J.C.; Beecham, G.W.; Grenier-Boley, B.; et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat. Genet. 2013, 45, 1452–1458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Miyamoto, A.; Wake, H.; Ishikawa, A.W.; Eto, K.; Shibata, K.; Murakoshi, H.; Koizumi, S.; Moorhouse, A.J.; Yoshimura, Y.; Nabekura, J. Microglia contact induces synapse formation in developing somatosensory cortex. Nat. Commun. 2016, 7, 12540. [Google Scholar] [CrossRef] [Green Version]
  85. Weinhard, L.; di Bartolomei, G.; Bolasco, G.; Machado, P.; Schieber, N.L.; Neniskyte, U.; Exiga, M.; Vadisiute, A.; Raggioli, A.; Schertel, A.; et al. Microglia remodel synapses by presynaptic trogocytosis and spine head filopodia induction. Nat. Commun. 2018, 9, 1228. [Google Scholar] [CrossRef] [Green Version]
  86. Antonucci, F.; Turola, E.; Riganti, L.; Caleo, M.; Gabrielli, M.; Perrotta, C.; Novellino, L.; Clementi, E.; Giussani, P.; Viani, P.; et al. Microvesicles released from microglia stimulate synaptic activity via enhanced sphingolipid metabolism. EMBO J. 2012, 31, 1231–1240. [Google Scholar] [CrossRef]
  87. Riganti, L.; Antonucci, F.; Gabrielli, M.; Prada, I.; Giussani, P.; Viani, P.; Valtorta, F.; Menna, E.; Matteoli, M.; Verderio, C. Sphingosine-1-Phosphate (S1P) Impacts Presynaptic Functions by Regulating Synapsin I Localization in the Presynaptic Compartment. J. Neurosci. 2016, 36, 4624–4634. [Google Scholar] [CrossRef] [Green Version]
  88. Marrone, M.C.; Morabito, A.; Giustizieri, M.; Chiurchiu, V.; Leuti, A.; Mattioli, M.; Marinelli, S.; Riganti, L.; Lombardi, M.; Murana, E.; et al. TRPV1 channels are critical brain inflammation detectors and neuropathic pain biomarkers in mice. Nat. Commun. 2017, 8, 15292. [Google Scholar] [CrossRef] [Green Version]
  89. Gabrielli, M.; Battista, N.; Riganti, L.; Prada, I.; Antonucci, F.; Cantone, L.; Matteoli, M.; Maccarrone, M.; Verderio, C. Active endocannabinoids are secreted on extracellular membrane vesicles. EMBO Rep. 2015, 16, 213–220. [Google Scholar] [CrossRef] [Green Version]
  90. Prada, I.; Gabrielli, M.; Turola, E.; Iorio, A.; D’Arrigo, G.; Parolisi, R.; De Luca, M.; Pacifici, M.; Bastoni, M.; Lombardi, M.; et al. Glia-to-neuron transfer of miRNAs via extracellular vesicles: A new mechanism underlying inflammation-induced synaptic alterations. Acta Neuropathol. 2018, 135, 529–550. [Google Scholar] [CrossRef]
  91. Vinuesa, A.; Bentivegna, M.; Calfa, G.; Filipello, F.; Pomilio, C.; Bonaventura, M.M.; Lux-Lantos, V.; Matzkin, M.E.; Gregosa, A.; Presa, J.; et al. Early Exposure to a High-Fat Diet Impacts on Hippocampal Plasticity: Implication of Microglia-Derived Exosome-like Extracellular Vesicles. Mol. Neurobiol. 2019, 56, 5075–5094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Joshi, P.; Benussi, L.; Furlan, R.; Ghidoni, R.; Verderio, C. Extracellular vesicles in Alzheimer’s disease: Friends or foes? Focus on abeta-vesicle interaction. Int. J. Mol. Sci. 2015, 16, 4800–4813. [Google Scholar] [CrossRef] [PubMed]
  93. Simon, D.; Garcia-Garcia, E.; Royo, F.; Falcon-Perez, J.M.; Avila, J. Proteostasis of tau. Tau overexpression results in its secretion via membrane vesicles. FEBS Lett. 2012, 586, 47–54. [Google Scholar] [CrossRef] [PubMed]
  94. Clayton, K.A.; Van Enoo, A.A.; Ikezu, T. Alzheimer’s Disease: The Role of Microglia in Brain Homeostasis and Proteopathy. Front. Neurosci. 2017, 11, 680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Gallart-Palau, X.; Guo, X.; Serra, A.; Sze, S.K. Alzheimer’s disease progression characterized by alterations in the molecular profiles and biogenesis of brain extracellular vesicles. Alzheimer’s Res. Ther. 2020, 12, 54. [Google Scholar] [CrossRef]
  96. D’Acunzo, P.; Hargash, T.; Pawlik, M.; Goulbourne, C.N.; Perez-Gonzalez, R.; Levy, E. Enhanced generation of intraluminal vesicles in neuronal late endosomes in the brain of a Down syndrome mouse model with endosomal dysfunction. Dev. Neurobiol. 2019, 79, 656–663. [Google Scholar] [CrossRef]
  97. Gauthier, S.A.; Perez-Gonzalez, R.; Sharma, A.; Huang, F.K.; Alldred, M.J.; Pawlik, M.; Kaur, G.; Ginsberg, S.D.; Neubert, T.A.; Levy, E. Enhanced exosome secretion in Down syndrome brain—A protective mechanism to alleviate neuronal endosomal abnormalities. Acta Neuropathol. Commun. 2017, 5, 65. [Google Scholar] [CrossRef] [Green Version]
  98. Agosta, F.; Dalla Libera, D.; Spinelli, E.G.; Finardi, A.; Canu, E.; Bergami, A.; Bocchio Chiavetto, L.; Baronio, M.; Comi, G.; Martino, G.; et al. Myeloid microvesicles in cerebrospinal fluid are associated with myelin damage and neuronal loss in mild cognitive impairment and Alzheimer disease. Ann. Neurol. 2014, 76, 813–825. [Google Scholar] [CrossRef]
  99. Gabrielli, M.; Raffaele, S.; Fumagalli, M.; Verderio, C. The multiple faces of extracellular vesicles released by microglia: Where are we 10 years after? Front. Cell Neurosci. 2022, 16, 984690. [Google Scholar] [CrossRef]
  100. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef]
  101. Halassa, M.M.; Fellin, T.; Haydon, P.G. Tripartite synapses: Roles for astrocytic purines in the control of synaptic physiology and behavior. Neuropharmacology 2009, 57, 343–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. D’Arrigo, G.; Gabrielli, M.; Scaroni, F.; Swuec, P.; Amin, L.; Pegoraro, A.; Adinolfi, E.; Di Virgilio, F.; Cojoc, D.; Legname, G.; et al. Astrocytes-derived extracellular vesicles in motion at the neuron surface: Involvement of the prion protein. J. Extracell. Vesicles 2021, 10, e12114. [Google Scholar] [CrossRef] [PubMed]
  103. Venturini, A.; Passalacqua, M.; Pelassa, S.; Pastorino, F.; Tedesco, M.; Cortese, K.; Gagliani, M.C.; Leo, G.; Maura, G.; Guidolin, D.; et al. Exosomes From Astrocyte Processes: Signaling to Neurons. Front. Pharmacol. 2019, 10, 1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Datta Chaudhuri, A.; Dasgheyb, R.M.; DeVine, L.R.; Bi, H.; Cole, R.N.; Haughey, N.J. Stimulus-dependent modifications in astrocyte-derived extracellular vesicle cargo regulate neuronal excitability. Glia 2020, 68, 128–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Patel, M.R.; Weaver, A.M. Astrocyte-derived small extracellular vesicles promote synapse formation via fibulin-2-mediated TGF-beta signaling. Cell. Rep. 2021, 34, 108829. [Google Scholar] [CrossRef]
  106. You, Y.; Borgmann, K.; Edara, V.V.; Stacy, S.; Ghorpade, A.; Ikezu, T. Activated human astrocyte-derived extracellular vesicles modulate neuronal uptake, differentiation and firing. J. Extracell. Vesicles 2020, 9, 1706801. [Google Scholar] [CrossRef]
  107. Carter, S.F.; Scholl, M.; Almkvist, O.; Wall, A.; Engler, H.; Langstrom, B.; Nordberg, A. Evidence for astrocytosis in prodromal Alzheimer disease provided by 11C-deuterium-L-deprenyl: A multitracer PET paradigm combining 11C-Pittsburgh compound B and 18F-FDG. J. Nucl. Med. 2012, 53, 37–46. [Google Scholar] [CrossRef] [Green Version]
  108. Zhou, B.; Zuo, Y.X.; Jiang, R.T. Astrocyte morphology: Diversity, plasticity, and role in neurological diseases. CNS Neurosci. Ther. 2019, 25, 665–673. [Google Scholar] [CrossRef]
  109. St-Pierre, M.K.; Vander Zwaag, J.; Loewen, S.; Tremblay, M.E. All roads lead to heterogeneity: The complex involvement of astrocytes and microglia in the pathogenesis of Alzheimer’s disease. Front. Cell. Neurosci. 2022, 16, 932572. [Google Scholar] [CrossRef]
  110. Sadick, J.S.; O’Dea, M.R.; Hasel, P.; Dykstra, T.; Faustin, A.; Liddelow, S.A. Astrocytes and oligodendrocytes undergo subtype-specific transcriptional changes in Alzheimer’s disease. Neuron 2022, 110, 1788–1805.e1710. [Google Scholar] [CrossRef]
  111. Sidoryk-Wegrzynowicz, M.; Gerber, Y.N.; Ries, M.; Sastre, M.; Tolkovsky, A.M.; Spillantini, M.G. Astrocytes in mouse models of tauopathies acquire early deficits and lose neurosupportive functions. Acta Neuropathol. Commun. 2017, 5, 89. [Google Scholar] [CrossRef] [PubMed]
  112. Richetin, K.; Steullet, P.; Pachoud, M.; Perbet, R.; Parietti, E.; Maheswaran, M.; Eddarkaoui, S.; Begard, S.; Pythoud, C.; Rey, M.; et al. Tau accumulation in astrocytes of the dentate gyrus induces neuronal dysfunction and memory deficits in Alzheimer’s disease. Nat. Neurosci. 2020, 23, 1567–1579. [Google Scholar] [CrossRef]
  113. Jones, R.S.; Minogue, A.M.; Connor, T.J.; Lynch, M.A. Amyloid-beta-induced astrocytic phagocytosis is mediated by CD36, CD47 and RAGE. J. Neuroimmune Pharmacol. 2013, 8, 301–311. [Google Scholar] [CrossRef]
  114. Chung, W.S.; Clarke, L.E.; Wang, G.X.; Stafford, B.K.; Sher, A.; Chakraborty, C.; Joung, J.; Foo, L.C.; Thompson, A.; Chen, C.; et al. Astrocytes mediate synapse elimination through MEGF10 and MERTK pathways. Nature 2013, 504, 394–400. [Google Scholar] [CrossRef] [Green Version]
  115. Chang, G.H.; Barbaro, N.M.; Pieper, R.O. Phosphatidylserine-dependent phagocytosis of apoptotic glioma cells by normal human microglia, astrocytes, and glioma cells. Neuro Oncol. 2000, 2, 174–183. [Google Scholar] [CrossRef]
  116. Magnus, T.; Chan, A.; Linker, R.A.; Toyka, K.V.; Gold, R. Astrocytes are less efficient in the removal of apoptotic lymphocytes than microglia cells: Implications for the role of glial cells in the inflamed central nervous system. J. Neuropathol. Exp. Neurol. 2002, 61, 760–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sokolowski, J.D.; Nobles, S.L.; Heffron, D.S.; Park, D.; Ravichandran, K.S.; Mandell, J.W. Brain-specific angiogenesis inhibitor-1 expression in astrocytes and neurons: Implications for its dual function as an apoptotic engulfment receptor. Brain Behav. Immun. 2011, 25, 915–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Loov, C.; Hillered, L.; Ebendal, T.; Erlandsson, A. Engulfing astrocytes protect neurons from contact-induced apoptosis following injury. PLoS ONE 2012, 7, e33090. [Google Scholar] [CrossRef] [Green Version]
  119. Nielsen, H.M.; Mulder, S.D.; Belien, J.A.; Musters, R.J.; Eikelenboom, P.; Veerhuis, R. Astrocytic A beta 1-42 uptake is determined by A beta-aggregation state and the presence of amyloid-associated proteins. Glia 2010, 58, 1235–1246. [Google Scholar] [CrossRef]
  120. Nagele, R.G.; D’Andrea, M.R.; Lee, H.; Venkataraman, V.; Wang, H.Y. Astrocytes accumulate A beta 42 and give rise to astrocytic amyloid plaques in Alzheimer disease brains. Brain Res. 2003, 971, 197–209. [Google Scholar] [CrossRef]
  121. Loov, C.; Mitchell, C.H.; Simonsson, M.; Erlandsson, A. Slow degradation in phagocytic astrocytes can be enhanced by lysosomal acidification. Glia 2015, 63, 1997–2009. [Google Scholar] [CrossRef] [PubMed]
  122. Sollvander, S.; Nikitidou, E.; Brolin, R.; Soderberg, L.; Sehlin, D.; Lannfelt, L.; Erlandsson, A. Accumulation of amyloid-beta by astrocytes result in enlarged endosomes and microvesicle-induced apoptosis of neurons. Mol. Neurodegener. 2016, 11, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Nikitidou, E.; Khoonsari, P.E.; Shevchenko, G.; Ingelsson, M.; Kultima, K.; Erlandsson, A. Increased Release of Apolipoprotein E in Extracellular Vesicles Following Amyloid-beta Protofibril Exposure of Neuroglial Co-Cultures. J. Alzheimer’s Dis. 2017, 60, 305–321. [Google Scholar] [CrossRef] [Green Version]
  124. Gonzalez-Molina, L.A.; Villar-Vesga, J.; Henao-Restrepo, J.; Villegas, A.; Lopera, F.; Cardona-Gomez, G.P.; Posada-Duque, R. Extracellular Vesicles from 3xTg-AD Mouse and Alzheimer’s Disease Patient Astrocytes Impair Neuroglial and Vascular Components. Front. Aging Neurosci. 2021, 13, 593927. [Google Scholar] [CrossRef]
  125. Dinkins, M.B.; Dasgupta, S.; Wang, G.; Zhu, G.; Bieberich, E. Exosome reduction in vivo is associated with lower amyloid plaque load in the 5XFAD mouse model of Alzheimer’s disease. Neurobiol. Aging 2014, 35, 1792–1800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Goetzl, E.J.; Schwartz, J.B.; Abner, E.L.; Jicha, G.A.; Kapogiannis, D. High complement levels in astrocyte-derived exosomes of Alzheimer disease. Ann. Neurol. 2018, 83, 544–552. [Google Scholar] [CrossRef]
  127. Muraoka, S.; Jedrychowski, M.P.; Yanamandra, K.; Ikezu, S.; Gygi, S.P.; Ikezu, T. Proteomic Profiling of Extracellular Vesicles Derived from Cerebrospinal Fluid of Alzheimer’s Disease Patients: A Pilot Study. Cells 2020, 9, 1959. [Google Scholar] [CrossRef]
  128. You, Y.; Muraoka, S.; Jedrychowski, M.P.; Hu, J.; McQuade, A.K.; Young-Pearse, T.; Aslebagh, R.; Shaffer, S.A.; Gygi, S.P.; Blurton-Jones, M.; et al. Human neural cell type-specific extracellular vesicle proteome defines disease-related molecules associated with activated astrocytes in Alzheimer’s disease brain. J. Extracell. Vesicles 2022, 11, e12183. [Google Scholar] [CrossRef]
  129. Stronati, E.; Conti, R.; Cacci, E.; Cardarelli, S.; Biagioni, S.; Poiana, G. Extracellular Vesicle-Induced Differentiation of Neural Stem Progenitor Cells. Int. J. Mol. Sci. 2019, 20, 3691. [Google Scholar] [CrossRef] [Green Version]
  130. Ma, Y.; Li, C.; Huang, Y.; Wang, Y.; Xia, X.; Zheng, J.C. Exosomes released from neural progenitor cells and induced neural progenitor cells regulate neurogenesis through miR-21a. Cell. Commun. Signal. 2019, 17, 96. [Google Scholar] [CrossRef]
  131. Budnik, V.; Ruiz-Canada, C.; Wendler, F. Extracellular vesicles round off communication in the nervous system. Nat. Rev. Neurosci. 2016, 17, 160–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Kramer-Albers, E.M.; Hill, A.F. Extracellular vesicles: Interneural shuttles of complex messages. Curr. Opin. Neurobiol. 2016, 39, 101–107. [Google Scholar] [CrossRef] [PubMed]
  133. Bahrini, I.; Song, J.H.; Diez, D.; Hanayama, R. Neuronal exosomes facilitate synaptic pruning by up-regulating complement factors in microglia. Sci. Rep. 2015, 5, 7989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Morel, L.; Regan, M.; Higashimori, H.; Ng, S.K.; Esau, C.; Vidensky, S.; Rothstein, J.; Yang, Y. Neuronal exosomal miRNA-dependent translational regulation of astroglial glutamate transporter GLT1. J. Biol. Chem. 2013, 288, 7105–7116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chivet, M.; Javalet, C.; Laulagnier, K.; Blot, B.; Hemming, F.J.; Sadoul, R. Exosomes secreted by cortical neurons upon glutamatergic synapse activation specifically interact with neurons. J. Extracell. Vesicles 2014, 3, 24722. [Google Scholar] [CrossRef] [Green Version]
  136. Lachenal, G.; Pernet-Gallay, K.; Chivet, M.; Hemming, F.J.; Belly, A.; Bodon, G.; Blot, B.; Haase, G.; Goldberg, Y.; Sadoul, R. Release of exosomes from differentiated neurons and its regulation by synaptic glutamatergic activity. Mol. Cell. Neurosci. 2011, 46, 409–418. [Google Scholar] [CrossRef] [Green Version]
  137. Chivet, M.; Javalet, C.; Hemming, F.; Pernet-Gallay, K.; Laulagnier, K.; Fraboulet, S.; Sadoul, R. Exosomes as a novel way of interneuronal communication. Biochem. Soc. Trans. 2013, 41, 241–244. [Google Scholar] [CrossRef] [Green Version]
  138. Chivet, M.; Hemming, F.; Pernet-Gallay, K.; Fraboulet, S.; Sadoul, R. Emerging role of neuronal exosomes in the central nervous system. Front. Physiol. 2012, 3, 145. [Google Scholar] [CrossRef] [Green Version]
  139. Faure, J.; Lachenal, G.; Court, M.; Hirrlinger, J.; Chatellard-Causse, C.; Blot, B.; Grange, J.; Schoehn, G.; Goldberg, Y.; Boyer, V.; et al. Exosomes are released by cultured cortical neurones. Mol. Cell. Neurosci. 2006, 31, 642–648. [Google Scholar] [CrossRef]
  140. Brenna, S.; Altmeppen, H.C.; Mohammadi, B.; Rissiek, B.; Schlink, F.; Ludewig, P.; Krisp, C.; Schluter, H.; Failla, A.V.; Schneider, C.; et al. Characterization of brain-derived extracellular vesicles reveals changes in cellular origin after stroke and enrichment of the prion protein with a potential role in cellular uptake. J. Extracell. Vesicles 2020, 9, 1809065. [Google Scholar] [CrossRef]
  141. Rajendran, L.; Honsho, M.; Zahn, T.R.; Keller, P.; Geiger, K.D.; Verkade, P.; Simons, K. Alzheimer’s disease beta-amyloid peptides are released in association with exosomes. Proc. Natl. Acad. Sci. USA 2006, 103, 11172–11177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Sharples, R.A.; Vella, L.J.; Nisbet, R.M.; Naylor, R.; Perez, K.; Barnham, K.J.; Masters, C.L.; Hill, A.F. Inhibition of gamma-secretase causes increased secretion of amyloid precursor protein C-terminal fragments in association with exosomes. FASEB J. 2008, 22, 1469–1478. [Google Scholar] [CrossRef] [PubMed]
  143. Vingtdeux, V.; Hamdane, M.; Loyens, A.; Gele, P.; Drobeck, H.; Begard, S.; Galas, M.C.; Delacourte, A.; Beauvillain, J.C.; Buee, L.; et al. Alkalizing drugs induce accumulation of amyloid precursor protein by-products in luminal vesicles of multivesicular bodies. J. Biol. Chem. 2007, 282, 18197–18205. [Google Scholar] [CrossRef] [Green Version]
  144. Ghidoni, R.; Paterlini, A.; Albertini, V.; Glionna, M.; Monti, E.; Schiaffonati, L.; Benussi, L.; Levy, E.; Binetti, G. Cystatin C is released in association with exosomes: A new tool of neuronal communication which is unbalanced in Alzheimer’s disease. Neurobiol. Aging 2011, 32, 1435–1442. [Google Scholar] [CrossRef] [Green Version]
  145. Laulagnier, K.; Javalet, C.; Hemming, F.J.; Chivet, M.; Lachenal, G.; Blot, B.; Chatellard, C.; Sadoul, R. Amyloid precursor protein products concentrate in a subset of exosomes specifically endocytosed by neurons. Cell Mol. Life Sci. 2018, 75, 757–773. [Google Scholar] [CrossRef]
  146. Lauren, J.; Gimbel, D.A.; Nygaard, H.B.; Gilbert, J.W.; Strittmatter, S.M. Cellular prion protein mediates impairment of synaptic plasticity by amyloid-beta oligomers. Nature 2009, 457, 1128–1132. [Google Scholar] [CrossRef] [Green Version]
  147. 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] [PubMed] [Green Version]
  148. Yuyama, K.; Sun, H.; Igarashi, Y.; Monde, K.; Hirase, T.; Nakayama, M.; Makino, Y. Immuno-digital invasive cleavage assay for analyzing Alzheimer’s amyloid ss-bound extracellular vesicles. Alzheimer’s Res. Ther. 2022, 14, 140. [Google Scholar] [CrossRef] [PubMed]
  149. Yuyama, K.; Igarashi, Y. Linking glycosphingolipids to Alzheimer’s amyloid-ss: Extracellular vesicles and functional plant materials. Glycoconj. J. 2022, 39, 613–618. [Google Scholar] [CrossRef]
  150. Guix, F.X.; Corbett, G.T.; Cha, D.J.; Mustapic, M.; Liu, W.; Mengel, D.; Chen, Z.; Aikawa, E.; Young-Pearse, T.; Kapogiannis, D.; et al. Detection of Aggregation-Competent Tau in Neuron-Derived Extracellular Vesicles. Int. J. Mol. Sci. 2018, 19, 663. [Google Scholar] [CrossRef]
  151. Goetzl, E.J.; Abner, E.L.; Jicha, G.A.; Kapogiannis, D.; Schwartz, J.B. Declining levels of functionally specialized synaptic proteins in plasma neuronal exosomes with progression of Alzheimer’s disease. FASEB J. 2018, 32, 888–893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Goetzl, E.J.; Kapogiannis, D.; Schwartz, J.B.; Lobach, I.V.; Goetzl, L.; Abner, E.L.; Jicha, G.A.; Karydas, A.M.; Boxer, A.; Miller, B.L. Decreased synaptic proteins in neuronal exosomes of frontotemporal dementia and Alzheimer’s disease. FASEB J. 2016, 30, 4141–4148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Li, T.R.; Yao, Y.X.; Jiang, X.Y.; Dong, Q.Y.; Yu, X.F.; Wang, T.; Cai, Y.N.; Han, Y. beta-Amyloid in blood neuronal-derived extracellular vesicles is elevated in cognitively normal adults at risk of Alzheimer’s disease and predicts cerebral amyloidosis. Alzheimer’s Res. Ther. 2022, 14, 66. [Google Scholar] [CrossRef] [PubMed]
  154. Kapogiannis, D.; Mustapic, M.; Shardell, M.D.; Berkowitz, S.T.; Diehl, T.C.; Spangler, R.D.; Tran, J.; Lazaropoulos, M.P.; Chawla, S.; Gulyani, S.; et al. Association of Extracellular Vesicle Biomarkers with Alzheimer Disease in the Baltimore Longitudinal Study of Aging. JAMA Neurol. 2019, 76, 1340–1351. [Google Scholar] [CrossRef]
  155. Eren, E.; Hunt, J.F.V.; Shardell, M.; Chawla, S.; Tran, J.; Gu, J.; Vogt, N.M.; Johnson, S.C.; Bendlin, B.B.; Kapogiannis, D. Extracellular vesicle biomarkers of Alzheimer’s disease associated with sub-clinical cognitive decline in late middle age. Alzheimer’s Dement. 2020, 16, 1293–1304. [Google Scholar] [CrossRef]
  156. Tian, C.; Stewart, T.; Hong, Z.; Guo, Z.; Aro, P.; Soltys, D.; Pan, C.; Peskind, E.R.; Zabetian, C.P.; Shaw, L.M.; et al. Blood extracellular vesicles carrying synaptic function- and brain-related proteins as potential biomarkers for Alzheimer’s disease. Alzheimer’s Dement. 2022. [Google Scholar] [CrossRef]
  157. Perez-Gonzalez, R.; Kim, Y.; Miller, C.; Pacheco-Quinto, J.; Eckman, E.A.; Levy, E. Extracellular vesicles: Where the amyloid precursor protein carboxyl-terminal fragments accumulate and amyloid-beta oligomerizes. FASEB J. 2020, 34, 12922–12931. [Google Scholar] [CrossRef]
  158. Picciolini, S.; Gualerzi, A.; Carlomagno, C.; Cabinio, M.; Sorrentino, S.; Baglio, F.; Bedoni, M. An SPRi-based biosensor pilot study: Analysis of multiple circulating extracellular vesicles and hippocampal volume in Alzheimer’s disease. J. Pharm. Biomed. Anal. 2021, 192, 113649. [Google Scholar] [CrossRef]
  159. Zheng, T.; Pu, J.; Chen, Y.; Mao, Y.; Guo, Z.; Pan, H.; Zhang, L.; Zhang, H.; Sun, B.; Zhang, B. Plasma Exosomes Spread and Cluster Around beta-Amyloid Plaques in an Animal Model of Alzheimer’s Disease. Front. Aging Neurosci. 2017, 9, 12. [Google Scholar] [CrossRef] [Green Version]
  160. Verweij, F.J.; Balaj, L.; Boulanger, C.M.; Carter, D.R.F.; Compeer, E.B.; D’Angelo, G.; El Andaloussi, S.; Goetz, J.G.; Gross, J.C.; Hyenne, V.; et al. The power of imaging to understand extracellular vesicle biology in vivo. Nat. Methods 2021, 18, 1013–1026. [Google Scholar] [CrossRef]
  161. Kur, I.M.; Prouvot, P.H.; Fu, T.; Fan, W.; Muller-Braun, F.; Das, A.; Das, S.; Deller, T.; Roeper, J.; Stroh, A.; et al. Neuronal activity triggers uptake of hematopoietic extracellular vesicles in vivo. PLoS Biol. 2020, 18, e3000643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Neckles, V.N.; Morton, M.C.; Holmberg, J.C.; Sokolov, A.M.; Nottoli, T.; Liu, D.; Feliciano, D.M. A transgenic inducible GFP extracellular-vesicle reporter (TIGER) mouse illuminates neonatal cortical astrocytes as a source of immunomodulatory extracellular vesicles. Sci. Rep. 2019, 9, 3094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Heusermann, W.; Hean, J.; Trojer, D.; Steib, E.; von Bueren, S.; Graff-Meyer, A.; Genoud, C.; Martin, K.; Pizzato, N.; Voshol, J.; et al. Exosomes surf on filopodia to enter cells at endocytic hot spots, traffic within endosomes, and are targeted to the ER. J. Cell. Biol. 2016, 213, 173–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Prada, I.; Amin, L.; Furlan, R.; Legname, G.; Verderio, C.; Cojoc, D. A new approach to follow a single extracellular vesicle-cell interaction using optical tweezers. Biotechniques 2016, 60, 35–41. [Google Scholar] [CrossRef] [Green Version]
  165. Song, Y.; Li, Z.; He, T.; Qu, M.; Jiang, L.; Li, W.; Shi, X.; Pan, J.; Zhang, L.; Wang, Y.; et al. M2 microglia-derived exosomes protect the mouse brain from ischemia-reperfusion injury via exosomal miR-124. Theranostics 2019, 9, 2910–2923. [Google Scholar] [CrossRef]
  166. Fevrier, B.; Vilette, D.; Archer, F.; Loew, D.; Faigle, W.; Vidal, M.; Laude, H.; Raposo, G. Cells release prions in association with exosomes. Proc. Natl. Acad. Sci. USA 2004, 101, 9683–9688. [Google Scholar] [CrossRef] [Green Version]
  167. Lombardi, M.; Gabrielli, M.; Adinolfi, E.; Verderio, C. Role of ATP in Extracellular Vesicle Biogenesis and Dynamics. Front. Pharmacol. 2021, 12, 654023. [Google Scholar] [CrossRef]
  168. Tonoli, E.; Verduci, I.; Gabrielli, M.; Prada, I.; Forcaia, G.; Coveney, C.; Savoca, M.P.; Boocock, D.J.; Sancini, G.; Mazzanti, M.; et al. Extracellular transglutaminase-2, nude or associated with astrocytic extracellular vesicles, modulates neuronal calcium homeostasis. Prog. Neurobiol. 2022, 216, 102313. [Google Scholar] [CrossRef]
  169. Wilhelmus, M.M.M.; Tonoli, E.; Coveney, C.; Boocock, D.J.; Jongenelen, C.A.M.; Breve, J.J.P.; Verderio, E.A.M.; Drukarch, B. The Transglutaminase-2 Interactome in the APP23 Mouse Model of Alzheimer’s Disease. Cells 2022, 11, 389. [Google Scholar] [CrossRef]
  170. Williams, C.; Pazos, R.; Royo, F.; Gonzalez, E.; Roura-Ferrer, M.; Martinez, A.; Gamiz, J.; Reichardt, N.C.; Falcon-Perez, J.M. Assessing the role of surface glycans of extracellular vesicles on cellular uptake. Sci. Rep. 2019, 9, 11920. [Google Scholar] [CrossRef]
  171. Christianson, H.C.; Svensson, K.J.; van Kuppevelt, T.H.; Li, J.P.; Belting, M. Cancer cell exosomes depend on cell-surface heparan sulfate proteoglycans for their internalization and functional activity. Proc. Natl. Acad. Sci. USA 2013, 110, 17380–17385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Barres, C.; Blanc, L.; Bette-Bobillo, P.; Andre, S.; Mamoun, R.; Gabius, H.J.; Vidal, M. Galectin-5 is bound onto the surface of rat reticulocyte exosomes and modulates vesicle uptake by macrophages. Blood 2010, 115, 696–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Morelli, A.E.; Larregina, A.T.; Shufesky, W.J.; Sullivan, M.L.; Stolz, D.B.; Papworth, G.D.; Zahorchak, A.F.; Logar, A.J.; Wang, Z.; Watkins, S.C.; et al. Endocytosis, intracellular sorting, and processing of exosomes by dendritic cells. Blood 2004, 104, 3257–3266. [Google Scholar] [CrossRef] [Green Version]
  174. Sung, B.H.; Ketova, T.; Hoshino, D.; Zijlstra, A.; Weaver, A.M. Directional cell movement through tissues is controlled by exosome secretion. Nat. Commun. 2015, 6, 7164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Purushothaman, A.; Bandari, S.K.; Liu, J.; Mobley, J.A.; Brown, E.E.; Sanderson, R.D. Fibronectin on the Surface of Myeloma Cell-derived Exosomes Mediates Exosome-Cell Interactions. J. Biol. Chem. 2016, 291, 1652–1663. [Google Scholar] [CrossRef] [Green Version]
  176. van Niel, G.; D’Angelo, G.; Raposo, G. Shedding light on the cell biology of extracellular vesicles. Nat. Rev. Mol. Cell. Biol. 2018, 19, 213–228. [Google Scholar] [CrossRef]
  177. Bianco, F.; Pravettoni, E.; Colombo, A.; Schenk, U.; Moller, T.; Matteoli, M.; Verderio, C. Astrocyte-derived ATP induces vesicle shedding and IL-1 beta release from microglia. J. Immunol. 2005, 174, 7268–7277. [Google Scholar] [CrossRef] [Green Version]
  178. Hurwitz, S.N.; Sun, L.; Cole, K.Y.; Ford, C.R., 3rd; Olcese, J.M.; Meckes, D.G., Jr. An optimized method for enrichment of whole brain-derived extracellular vesicles reveals insight into neurodegenerative processes in a mouse model of Alzheimer’s disease. J. Neurosci. Methods 2018, 307, 210–220. [Google Scholar] [CrossRef]
  179. Alim, M.A.; Ma, Q.L.; Takeda, K.; Aizawa, T.; Matsubara, M.; Nakamura, M.; Asada, A.; Saito, T.; Kaji, H.; Yoshii, M.; et al. Demonstration of a role for alpha-synuclein as a functional microtubule-associated protein. J. Alzheimer’s Dis. 2004, 6, 435–442; discussion 443-439. [Google Scholar] [CrossRef]
  180. Westergard, L.; Christensen, H.M.; Harris, D.A. The cellular prion protein (PrP(C)): Its physiological function and role in disease. Biochim. Biophys. Acta 2007, 1772, 629–644. [Google Scholar] [CrossRef]
  181. Deyts, C.; Thinakaran, G.; Parent, A.T. APP Receptor? To Be or Not to Be. Trends Pharmacol. Sci. 2016, 37, 390–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Sabo, S.L.; Ikin, A.F.; Buxbaum, J.D.; Greengard, P. The amyloid precursor protein and its regulatory protein, FE65, in growth cones and synapses in vitro and in vivo. J. Neurosci. 2003, 23, 5407–5415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Brody, A.H.; Strittmatter, S.M. Synaptotoxic Signaling by Amyloid Beta Oligomers in Alzheimer’s Disease through Prion Protein and mGluR5. Adv. Pharmacol. 2018, 82, 293–323. [Google Scholar] [CrossRef] [PubMed]
  184. Nieznanski, K.; Surewicz, K.; Chen, S.; Nieznanska, H.; Surewicz, W.K. Interaction between prion protein and Abeta amyloid fibrils revisited. ACS Chem. Neurosci. 2014, 5, 340–345. [Google Scholar] [CrossRef]
  185. Corbett, G.T.; Wang, Z.; Hong, W.; Colom-Cadena, M.; Rose, J.; Liao, M.; Asfaw, A.; Hall, T.C.; Ding, L.; DeSousa, A.; et al. PrP is a central player in toxicity mediated by soluble aggregates of neurodegeneration-causing proteins. Acta Neuropathol. 2020, 139, 503–526. [Google Scholar] [CrossRef] [Green Version]
  186. Gulisano, W.; Maugeri, D.; Baltrons, M.A.; Fa, M.; Amato, A.; Palmeri, A.; D’Adamio, L.; Grassi, C.; Devanand, D.P.; Honig, L.S.; et al. Role of Amyloid-beta and Tau Proteins in Alzheimer’s Disease: Confuting the Amyloid Cascade. J. Alzheimer’s Dis. 2018, 64, S611–S631. [Google Scholar] [CrossRef]
  187. Shimoda, M.; Khokha, R. Metalloproteinases in extracellular vesicles. Biochim. Biophys. Acta Mol. Cell. Res. 2017, 1864, 1989–2000. [Google Scholar] [CrossRef]
  188. Lewin, S.; Hunt, S.; Lambert, D.W. Extracellular vesicles and the extracellular matrix: A new paradigm or old news? Biochem. Soc. Trans. 2020, 48, 2335–2345. [Google Scholar] [CrossRef]
  189. Rilla, K.; Mustonen, A.M.; Arasu, U.T.; Harkonen, K.; Matilainen, J.; Nieminen, P. Extracellular vesicles are integral and functional components of the extracellular matrix. Matrix. Biol. 2019, 75–76, 201–219. [Google Scholar] [CrossRef]
  190. Rackov, G.; Garcia-Romero, N.; Esteban-Rubio, S.; Carrion-Navarro, J.; Belda-Iniesta, C.; Ayuso-Sacido, A. Vesicle-Mediated Control of Cell Function: The Role of Extracellular Matrix and Microenvironment. Front. Physiol. 2018, 9, 651. [Google Scholar] [CrossRef]
  191. Buzas, E.I.; Toth, E.A.; Sodar, B.W.; Szabo-Taylor, K.E. Molecular interactions at the surface of extracellular vesicles. Semin. Immunopathol. 2018, 40, 453–464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Sethi, M.K.; Zaia, J. Extracellular matrix proteomics in schizophrenia and Alzheimer’s disease. Anal. Bioanal. Chem. 2017, 409, 379–394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Kramer-Albers, E.M. Extracellular vesicles in the oligodendrocyte microenvironment. Neurosci. Lett. 2020, 725, 134915. [Google Scholar] [CrossRef]
  194. Chen, X.D.; Zhao, J.; Yang, X.; Zhou, B.W.; Yan, Z.; Liu, W.F.; Li, C.; Liu, K.X. Gut-Derived Exosomes Mediate Memory Impairment After Intestinal Ischemia/Reperfusion via Activating Microglia. Mol. Neurobiol. 2021, 58, 4828–4841. [Google Scholar] [CrossRef] [PubMed]
  195. Cambria, C.; Ingegnoli, F.; Borzi, E.; Cantone, L.; Coletto, L.A.; Rizzuto, A.S.; De Lucia, O.; Briguglio, S.; Ruscica, M.; Caporali, R.; et al. Synovial Fluid-Derived Extracellular Vesicles of Patients with Arthritides Contribute to Hippocampal Synaptic Dysfunctions and Increase with Mood Disorders Severity in Humans. Cells 2022, 11, 2276. [Google Scholar] [CrossRef] [PubMed]
  196. Vandendriessche, C.; Kapogiannis, D.; Vandenbroucke, R.E. Biomarker and therapeutic potential of peripheral extracellular vesicles in Alzheimer’s disease. Adv. Drug Deliv. Rev. 2022, 190, 114486. [Google Scholar] [CrossRef]
Figure 1. Effects of EVs released by brain cells on neuronal function in AD. Graphic summary of the effects of EVs from different brain cell types on neuronal function in AD pathology. Parts of the figures were drawn by using pictures from Servier Medical Art, licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/; access date 20 November 2022). Asai et al., 2015 [15], Ruan et al., 2020 [54], Clayton et al, 2021 [55], Zhu et al., 2022 [56], Mallach et al., 2021a, 2021b [57,58], Gabrielli et al., 2022 [17], Wang et al., 2017 [59], Sardar-Sinha et al., 2018 [26], An et al., 2013 [60], Yuyama et al., 2012, 2014, 2015 [61,62,63], Eitan et al., 2016 [25], Polanco et al., 2018 [64], Ruan et al., 2021 [18], Leroux et al., 2022 [65], Joshi et al., 2014 [21], Beretta et al., 2020 [66], Wang et al., 2012 [67], Nogueras-Ortiz et al., 2020 [68].
Figure 1. Effects of EVs released by brain cells on neuronal function in AD. Graphic summary of the effects of EVs from different brain cell types on neuronal function in AD pathology. Parts of the figures were drawn by using pictures from Servier Medical Art, licensed under a Creative Commons Attribution 3.0 Unported License (https://creativecommons.org/licenses/by/3.0/; access date 20 November 2022). Asai et al., 2015 [15], Ruan et al., 2020 [54], Clayton et al, 2021 [55], Zhu et al., 2022 [56], Mallach et al., 2021a, 2021b [57,58], Gabrielli et al., 2022 [17], Wang et al., 2017 [59], Sardar-Sinha et al., 2018 [26], An et al., 2013 [60], Yuyama et al., 2012, 2014, 2015 [61,62,63], Eitan et al., 2016 [25], Polanco et al., 2018 [64], Ruan et al., 2021 [18], Leroux et al., 2022 [65], Joshi et al., 2014 [21], Beretta et al., 2020 [66], Wang et al., 2012 [67], Nogueras-Ortiz et al., 2020 [68].
Cells 12 00063 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gabrielli, M.; Tozzi, F.; Verderio, C.; Origlia, N. Emerging Roles of Extracellular Vesicles in Alzheimer’s Disease: Focus on Synaptic Dysfunction and Vesicle–Neuron Interaction. Cells 2023, 12, 63. https://doi.org/10.3390/cells12010063

AMA Style

Gabrielli M, Tozzi F, Verderio C, Origlia N. Emerging Roles of Extracellular Vesicles in Alzheimer’s Disease: Focus on Synaptic Dysfunction and Vesicle–Neuron Interaction. Cells. 2023; 12(1):63. https://doi.org/10.3390/cells12010063

Chicago/Turabian Style

Gabrielli, Martina, Francesca Tozzi, Claudia Verderio, and Nicola Origlia. 2023. "Emerging Roles of Extracellular Vesicles in Alzheimer’s Disease: Focus on Synaptic Dysfunction and Vesicle–Neuron Interaction" Cells 12, no. 1: 63. https://doi.org/10.3390/cells12010063

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