Next Article in Journal
Differences in MPS I and MPS II Disease Manifestations
Next Article in Special Issue
JI017 Attenuates Oxaliplatin-Induced Cold Allodynia via Spinal TRPV1 and Astrocytes Inhibition in Mice
Previous Article in Journal
Cold Atmospheric Plasma Changes the Amino Acid Composition of Solutions and Influences the Anti-Tumor Effect on Melanoma Cells
Previous Article in Special Issue
Protective Effect of Memantine on Bergmann Glia and Purkinje Cells Morphology in Optogenetic Model of Neurodegeneration in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dysregulation of Astrocyte–Neuronal Communication in Alzheimer’s Disease

Department of Neuroscience, University of Minnesota, Minneapolis, MN 55455, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(15), 7887; https://doi.org/10.3390/ijms22157887
Submission received: 23 June 2021 / Revised: 19 July 2021 / Accepted: 20 July 2021 / Published: 23 July 2021

Abstract

:
Recent studies implicate astrocytes in Alzheimer’s disease (AD); however, their role in pathogenesis is poorly understood. Astrocytes have well-established functions in supportive functions such as extracellular ionic homeostasis, structural support, and neurovascular coupling. However, emerging research on astrocytic function in the healthy brain also indicates their role in regulating synaptic plasticity and neuronal excitability via the release of neuroactive substances named gliotransmitters. Here, we review how this “active” role of astrocytes at synapses could contribute to synaptic and neuronal network dysfunction and cognitive impairment in AD.

1. Introduction

The main independent variable linked to the risk of suffering Alzheimer’s disease (AD) is aging [1,2,3]. AD is the most common neurodegenerative disorder with the contribution of multiple cell types throughout the central nervous system. Most cases are sporadic, and only 2–5% of patients have a genetic familial background. Alois Alzheimer in 1907 first described the histopathological characteristics of the disease with the presence of senile plaques (formed by the extracellular aggregation of amyloid β (Aβ) protein) and neurofibrillary tangles (formed by the destructuring of the microtubules due to hyperphosphorylation of the tau protein) [4,5,6].
Even though the disease was described more than a century ago, its underlying causes are still unknown today. It is thought to be a disease of multifactorial etiology, with numerous risk factors, among them genetic, age, sex, educational level, and diet. The systematic biochemical investigation of the brains of patients with AD in the late 1960s and early 1970s led to the cholinergic hypothesis [7]. Thus, it was proposed that the degeneration of cholinergic neurons in the basal forebrain and the associated loss of cholinergic neurotransmission in the cerebral cortex and other areas contribute significantly to the deterioration in cognitive function seen in patients with AD [7,8]. Later, the observation of accumulation and deposition of oligomeric or fibrillary Aβ peptides in AD patient brains popularized the amyloidogenic hypothesis as the primary cause of AD and remains a mainstream concept [4,9,10]. The central role of the Aβ peptide in AD pathogenesis has been supported by the fact that familial forms of the disease are caused by its overproduction. The Aβ peptide is produced by processing of the amyloid-β protein precursor (APP) through sequential cleavage by β- and γ-secretases in the amyloidogenic pathway [10]. Indeed, the majority of the AD animal models used today are mice transgenic for human APP and Aβ [11]. Likewise, the contribution of microtubule-associated protein tau in AD pathology has also been proposed [12,13]. In this hypothesis, the presence of abnormally high levels of tau protein in the hyperphosphorylated state promotes the production of toxic oligomeric tau and paired helical filaments, which further assemble into toxic neurofibrillary tangles. Thus, it is clear that more studies are needed to explain the etiopathogenesis of AD, which is a critical step for developing rational treatments.
The most predominant and striking sign in an AD patient is the progressive decline in cognition, which is primarily due to the loss of neurons and synapses in the hippocampal formation and related areas [14]. As expected, research on AD has focused mainly on neuronal cells, which allowed great insights on the neurocentric molecular pathways involved in these processes, such as protein degradation systems, post-translational modifications, and interactions with cytoskeletons. However, in the last three decades, growing evidence suggests that glial cells, in particular astrocytes, have a contributing role in AD pathophysiology. Astrocytes in the human brain play key roles in numerous functions within the central nervous system such as structural support, ionic balance of the extracellular space, neurotransmitter clearance at synapses, and modulation of synaptic signaling. Thus, it is hardly surprising that astrocytes have been implicated in the pathology of several neurodegenerative diseases, including AD. Numerous excellent reviews have discussed the role of astrocytes in AD in the context of altered inflammatory processes [15,16,17], amyloid clearance [18], and neurovascular coupling [19,20]. The present review focuses instead on the role of dysregulated bidirectional communication of neurons and astrocytes in AD pathology.

2. Basic Structure and Function of Astrocytes

The term astrocyte was created to describe star-shaped glial cells detected in histological brain specimens [21]. Glia often was considered to vastly outnumber neurons in the brain but a re-examination of various counting methods place glia/neuron ratios closer to 1:1 in the human brain [22]. Traditionally, two major classes of astrocytes have been distinguished in histological sections of the central nervous system (CNS) based on their morphology and distribution [23]. The fibrous astrocytes are mainly in the white matter with few processes, and protoplasmic astrocytes are mostly found in gray matter and characterized by their intricate morphology with complex branching processes. More recently, RNA and proteomic analysis have identified several distinct astrocyte subtypes in diverse brain areas reflecting the distribution of morphologically and physiologically distinct astrocyte populations [24,25,26].
The tridimensional reconstruction of individual protoplasmic astrocytes in the rat hippocampus showed that astrocyte cell bodies are evenly spaced, and their processes overlap only minimally, creating a ‘tiling” of astrocytes [27,28]. This seems to be the case in other brain regions as well [29]. Furthermore, a single astrocyte in the rat hippocampus is estimated to occupy a territory of 66,000 µm3 of neuropil and contact over 140,000 synapses [27], placing a single astrocyte in a favorable position to modulate the activity of a vast neuronal network. Astrocytes in the human cortex have even larger territories and complex structural features [30]. Compared to rodents, the volume of human astrocytes is about 27-fold greater with a 2.5-fold increased diameter and each astrocyte is estimated to contact up to ≈2,000,000 synapses [30].
Traditionally, an important function attributed to astrocytes has been the removal or clearance of synaptic neurotransmitters such as glutamate or GABA [31,32]. Other “supporting” functions attributed to astrocytes comprise their participation in neural development, the buffering of extracellular potassium, and the regulation of blood flow [33,34,35,36,37].
Comparatively to neurons, the functional properties of astrocytes have received little attention. One major reason for this is the lack of active voltage-gated conductances in astrocytes. Electrophysiologically, astrocytes under voltage-clamped conditions display a quasi-linear voltage–current relationship with a cellular membrane almost exclusively permeable to potassium ions [38]. This potassium selective membrane is endowed by the expression of large amounts of inwardly rectifying potassium (Kir) channels, conferring astrocytes with their characteristic low input resistance and membrane potential close to the predicted equilibrium potential for transmembrane potassium. Molecular cloning of Kir channels in astrocytes shows that they are mainly weakly inwardly rectifying Kir4.1 channels [39,40]. Another major conductance found in astrocytes is connexin 43, which provides intercellular electrical coupling among astrocytes [41]. These electrophysiological features of astrocytes led to the hypothesis of “potassium spatial buffering”. In this hypothesis, when a localized release of potassium occurs due to enhanced neuronal activity, the excess extracellular potassium would enter the astrocytic syncytium and be expelled in other areas where the extracellular potassium is low [38].

3. Astrocytic Release of Gliotransmitters

Mainly due to the development of organic and genetically encoded calcium indicators and advances in imaging techniques, it has become clear that astrocytes have an “active” role in modulating the activity of neuronal circuits [42,43,44,45]. During synaptic activity, the release of neuronal transmitters leads to changes in intracellular calcium activity in astrocytes. This is so because astrocytes express a myriad of G-protein-coupled receptors (GPCRs) that respond to neurotransmitters through the activation of inositol triphosphate type 2 receptors (IP3R2), which mediate calcium release from the endoplasmic reticulum [46,47]. Such calcium elevations have been observed both in the astrocytic processes or microdomains and in their somata [37,48,49]. An outcome of the intracellular calcium changes in astrocytes is the release of neuroactive molecules or “gliotransmitters” [50]. Among them are ATP, glutamate, D-serine, and GABA [47,50,51,52,53,54,55,56,57,58,59,60]. Through the release of gliotransmitters, astrocytes have been found to modulate neuronal activity and synaptic transmission in several brain areas [47,55,61,62] and to impact animal behavior [63,64,65]. This bidirectional exchange of information between astrocytes and neurons is embodied in the concept of the tripartite synapse that includes astrocytes as integral elements of synaptic function along with presynaptic and postsynaptic processes (Figure 1) [42,44,47].
Remarkably, gliotransmitters modulate many forms of synaptic plasticity at various temporal and spatial scales [47,66]. Astrocytic modulation of presynaptic or postsynaptic neuronal terminals has been associated with short-term synaptic potentiation or depression [66]. Astrocytic modulation of long-lasting synaptic changes as long-term potentiation (LTP) and long-term depression (LTD) [47] is also extensively documented.

4. Reactive Astrocytes in Neurodegenerative Diseases

A hallmark sign of neurodegenerative disease is reactive gliosis with astrocytic morphological and functional changes [67,68,69]. Astrocyte reactivity is initially characterized by the hypertrophy of soma and processes triggered by inflammatory molecules [33,68]. In addition, there is an increased expression of intermediary filament proteins such as Glial Fibrillary Acidic Protein (GFAP) or vimentin [70,71]. Many studies have shown that in AD, there is a prominent presence of reactive astrocytes, in particular, surrounding amyloid plaques [72,73]. They are detected at the early phases of AD even before neuronal death and are found ubiquitously throughout disease progression [74]. In addition to morphological changes, they display graded reactivity with heterogeneous gene expression, morphology, and function [74,75]. Concerning this review, an important consequence of astrocyte reactivity in disease is the abnormal gliotransmitter release [76].

5. Altered Astrocyte Calcium Excitability in AD

There is great consensus among all available studies that in parallel with morphological changes and enhanced GFAP expression, spontaneous calcium signals in astrocytes are heightened in brains of AD animal models [77,78,79,80,81,82]. This increase appears to be independent of neuronal activity, since the blockade of neuronal activity with the sodium channel blocker tetrodotoxin does not instigate changes in spontaneous calcium activity [77,78,79]. This is in contrast to what is observed in normal aging, where no changes or even fewer spontaneous calcium oscillations occur [83,84].
More specifically, Kuchibhotla et al. found that astrocytes in the adult APP/PS1 mice (6 to 8 months old) with cortical amyloid plaques exhibited a significant increase in calcium transients with events synchronously coordinated across long distances and uncoupled from neuronal activity [78]. Likewise, Takano et al. using mice expressing the Swedish mutation of the APP gene showed that astrocytes exhibit a higher frequency of spontaneous oscillations even before the appearance of amyloid plaques [81]. Mechanistically, the increase in calcium activity in astrocytes may be mediated through the upregulation of neurotransmitter receptors in astrocytes such as the purinergic receptor P2Y1R [80,85], α7 nicotinic acetylcholine receptors (α7 nAChR), or metabotropic glutamate receptors mGluR5 [86,87]. Thus, the application of Aβ25–35 peptide, a neurotoxic Aβ fragment present in AD patients, upregulates calcium transients in primary cortical astrocytes, which can be blocked with P2 receptor antagonists [88]. Likewise, studies in vivo show that the enhanced calcium activity of astrocytes near Aβ plaques in APPPS1 mice is reduced after P2 purinoreceptor blockade [77].
Alternatively, Aβ plaques may bind and activate/inactivate α7 nAChR in astrocytes leading to hyperactivity in calcium signaling. In the human hippocampus and entorhinal cortex, α7, but not α4, subunit immunoreactivity is associated with astrocytes [86]. Moreover, a higher proportion of α7 expressing astrocytes was found in samples from AD patients compared with age-matched controls [86,87]. Evidence of an effect of Aβ plaques or oligomers on α7 nAChRs in astrocytes was obtained in hippocampal slices. Astrocytic calcium elevations can be elicited with nicotine and Aβ1-42 application in hippocampal slices [89]. These responses were present even in the presence of a sodium blocker, suggesting a neuronal-independent effect [89]. This scenario is supported by the fact that the incubation of rat hippocampal slices with Aβ1-42 at a concentration as low as 200 pM induces an increase in astrocytic calcium transients that is dependent on α7 nAChR expression or activity [90].
Given that homomeric α7 nAChR displays high permeability to calcium [91], stimulation of α7 nAChRs in astrocytes could lead to enhanced calcium activity in astrocytes via influx from extracellular sources. However, it has been shown that the increase in intracellular calcium in astrocytes via α7 nAChR signaling involves release from intracellular stores [92]. Since astrocytes can release gliotransmitters in response to intracellular calcium elevations, the upregulation of α7 nAChRs in astrocytes in AD may trigger the dysregulated release of gliotransmitters such as glutamate to ultimately induce neurotoxicity [79] (Figure 2).
Alternatively, the appearance of calcium hyperactivity in astrocytes may also be due to enhanced expression of the metabotropic glutamate receptor mGluR5 in astrocytes. Astrocytes detect glutamatergic transmission through mGluR5, and its activation regulates multiple forms of astrocyte–neuronal communication [93]. The relevance of mGluR5 signaling in astrocytes in adult brains remains somewhat controversial [94]; however, there seems a consensus of its upregulation in reactive astrocytes [95,96,97].
Thus, it is not surprising that the upregulation of mGluR5 is detected in astroglial cultures exposed to β-amyloid or in astrocytes in animal AD models and in postmortem human tissues [98,99,100,101,102]. Astrocytic mGluR5 activation by Aβ plaques leads to sustained calcium oscillations in the reactive astrocytes [103], which may trigger the release of intracellular glutamate, thus enhancing the neuronal excitability leading to excitotoxicity.
Several studies have also implicated the hyperactivity of calcium activity in astrocytes to the dysregulated function of neuronal networks in AD mouse models [104,105,106]. It has been hypothesized that this calcium hyperactivity of astrocytes networks may contribute to cognitive deficits observed in AD. Indeed, diminishing astrocyte calcium signaling pharmacologically or by astrocyte-specific genetic deletion (Ip3r2−/−) normalizes the neuronal network dysfunction and improves spatial memory in the APP/PS1 mouse model [80].
Notably, Richetin et al. found an accumulation of a tau isoform (3R) in hilar astrocytes of the dentate gyrus of patients afflicted with AD [107]. The overexpression of tau 3R isoform in dentate gyrus astrocytes in mice induced abnormal mitochondrial function in astrocytes and impaired synaptic and network activity in the hippocampus [107]. Behaviorally, tau accumulation in astrocytes also led to impairments in spatial memory tests. These results compellingly indicate that the accumulation of tau in astrocytes contributes to the pathogenesis of AD disease. So far, the impact of tau accumulation in astrocytic calcium excitability communication is unknown.

6. Altered Glutamate Release and Uptake from Astrocytes in AD

Glutamate is the major excitatory transmitter in the central nervous system and has a large array of physiological functions including learning and memory. Several families of glutamate receptor proteins have been identified: N-methyl-D-aspartate (NMDA) receptors, α-amino-3- hydroxy-5-methyl-4-isoxazole propionic acid (AMPA) receptors, kainate receptors, and metabotropic receptors [108].
Glutamate concentrations in the extracellular space are kept at low levels and tightly controlled by several mechanisms at the synapse. Perturbations to this regulatory system can lead to deleterious effects such as excess of extracellular glutamate, which can induce hyperexcitability in post-synaptic neurons to the point of excitotoxicity and cell death (cytotoxicity). It is usually accepted that the glutamate transporter activity of astrocytes is the primary mechanism responsible for clearing extracellular glutamate at synapses [109]. The major type of glutamate transporter expressed in astrocytes is the sodium-dependent symporter GLT-1 (GLT-1), since it is responsible for around 90% of astrocytic glutamate uptake in the brain [110], expressing at levels four or six times higher than the glutamate aspartate Transporter 1 (GLAST1) [111]. Moreover, the dysregulation of GLT-1 has been linked to neuronal cell death and neurological disorders, and the GLT-1 knockout mouse has a phenotype of lethal spontaneous seizures and significant neuronal loss [112].
Chronic glutamate excitotoxicity has been hypothesized to play a role in numerous neurodegenerative diseases including amyotrophic lateral sclerosis, AD, and Huntington’s disease [113]. The bulk of evidence indicates alterations in expression or subcellular localization of GLT-1 in AD [114,115,116,117]. Lower expression of GLT-1 mRNA in AD hippocampus and a 30% reduction of GLT-1 immunoreactivity in AD frontal cortex have been observed, with no concomitant decreases in GLAST [118]. Autopsied AD hippocampus showed decreased GLAST and GLT-1 mRNA expression with altered localization of the corresponding proteins into neurofibrillary tangles [119]. In line with the above-mentioned findings, functional studies indicate a reduction in glutamate uptake activity in AD brains. In autopsies of AD patients, with the decrease of the GLT immunoreactivity, there is also a decrease in glutamate transporter activity [118,120]. Reducing expression of GLT-1 in the AβPPswe/PS1ΔE9 mice accelerated cognitive deficits reminiscent of AD [121]. Accordingly, the overexpression of GLT-1 in the APP Swedish mice improved their cognitive functions and decreased pathology [122]. A more direct demonstration of altered glutamate homeostasis in AD mouse models is provided by Hefendehl et al. using in vivo two-photon imaging and the glutamate sensor iGluSnFR [123]. In the APPPS1 transgenic mice, the glutamate sensor shows that around Aβ plaques, there are high spontaneous and anomalous glutamate fluctuations [123]. Moreover, the dynamics of glutamate build-up and clearance upon sensory stimulation are also altered in these mice with a reduction of GLT-1 expression around Aβ plaques [123]. Conversely, AD patients with preserved cognitive function (no dementia) showed enhanced GLT-1 expression in comparison to controls (AD patients with dementia) [124]. Collectively, these studies demonstrate the potential importance of temporal changes of GLT-1 expression and localization in AD.
Astrocytes are not only involved in glutamate uptake via transporter activity, but they can also release glutamate as a gliotransmitter, most likely via a vesicular mechanism [125]. Glutamate released by astrocytes binds to either presynaptic metabotropic glutamate receptors or extrasynaptically-located postsynaptic NMDA receptors (NR2B) [126,127]. The latter evokes the so-called slow inward currents (SICs), which can be discriminated from the synaptic potentials by their much slower temporal kinetics [53,128,129,130,131]. The depolarizing action of these SICs is hypothesized to modulate neural excitability to affect neuronal action potential firing [132]. Moreover, because single astrocytes are near a large number (≈100) of neurons [29], SICs can be generated in many adjacent neurons to promote the synchrony of neuronal firing [53,133,134].
Overall, several studies report an increase in astrocytic-mediated SICs in animal or experimental models of AD [82,135,136]. For example, in rat hippocampal slices, application of the Aβ1-42 peptide not only induces increases in astrocytic calcium activity but also induces the increase of SIC frequency in CA1 pyramidal neurons [79] or cortical cultures [137] (Figure 2). Moreover, in the Tg2576 AD mouse model for Aβ over-production and accumulation, spontaneous astrocytic calcium elevations and SICs were also found [79]. More indirect measurements using fluorometric methods also show the effect of Aβ in eliciting glutamate release from astrocytes in hippocampal slices [135] or cortical cultures [88,136].
The heightened release of glutamate from astrocytes may underlie the changes in synaptic function and plasticity observed in animal models of AD. Activation of extrasynaptic NMDARs triggered by the glutamate release from astrocytes in response to Aβ peptide is followed rapidly by a decrease in miniature excitatory postsynaptic currents (mEPSC), representing initial synaptic dysfunction, and then by an increase in synaptic loss [137]. Ultimately, the combination of all the previous events may promote the cognitive impairments observed in behavioral tests [137].

7. Altered ATP Release from Astrocytes in AD

It is now well established that adenosine triphosphate (ATP) serves as an important neurotransmitter in the central and peripheral nervous system [138]. ATP mediates fast and slow synaptic potentials via ligand-gated cationic channels (P2X receptors) and G protein-coupled receptors (P2Y receptors). It is also now apparent that ATP is released from astrocytes as a gliotransmitter [139]. There has been a large body of evidence indicating multiple release mechanisms of ATP from astrocytes. Several studies have supported the notion that ATP is released from astrocytes in a calcium-dependent manner via exocytosis from synaptic-like vesicles [140,141,142,143,144]. Alternatively, several nonexocytotic release mechanisms also have been described for astrocytic ATP release. These include connexin hemichannels (connexin 43) [145], pannexin channels (Panx-1) [146], the calcium-dependent chloride channel bestrophin [141], and the calcium homeostasis modulator (CALHM) [147].
Enhancement of ATP release in hippocampal slices or astrocyte cultures is observed upon application of Aβ peptides [148]. In cultured rat astrocytes, exposure of astrocytes to Aβ1-42 increased the amplitude and velocity of evoked calcium waves mediated by enhanced ATP release [148]. Several studies have proposed that connexin 43 (Cx43) in a hemichannel function mediates the Aβ peptide induction of ATP release from astrocytes. Thus, Cx43 is upregulated in AD mouse models and AD human brains [149,150,151,152]. Moreover, amyloid exposure triggers the increased expression of Cx43 both in vitro and in vivo AD models [82,153,154]. Cx43 is also highly permeable to ATP [155], and knocking out the Cx43 gene in the APP/PPS1 mice reduces ATP release and decreases neuronal damage [140]. Finally, astrocytes of acute hippocampal slices containing Aβ plaques in the APP/PS1 mice show enhanced hemichannel activity and ATP release [82,88] (Figure 2).
ATP release from astrocytes in AD pathology may directly induce pathology. Conceivably, ATP activation of calcium-permeable pathways such as via P2X7 receptors or co-activation of NMDAR/P2X7 receptors could reduce the survival of neurons [82,135,136]. On the other hand, a protective role of ATP released from astrocytes in AD has also been suggested. Jung et al. found that exogenous ATP protects Aβ-mediated reduction of synaptic function in hippocampal neuronal cultures [156]. In addition, ATP prevented the Aβ-induced impairment of LTP in hippocampal slices [156].

8. Altered GABA Release from Astrocytes in AD

GABA is the major inhibitory neurotransmitter in the adult mammalian brain, and recent studies have also shown astrocytic release as a gliotransmitter in AD [157]. In the healthy brain, cytosolic concentrations in astrocytes are kept at low levels. However, in human AD patients and mouse models of AD, the astrocytic GABA cytosolic levels are unusually high [158]. Atypically, astrocytes seem to synthesize GABA from putrescine as a substrate via monoamine oxidation instead of using glutamate as a substrate [159]. GABA released from astrocytes exert a tonic inhibitory influence onto cerebellar granule neurons and striatal medium spiny neurons via the activation of GABAA receptors [159]. In the prefrontal cortex, on the other hand, astrocytic GABA interacts with GABAergic interneurons via GABAB receptors [160]. Under normal conditions, hippocampal astrocytes contain very little GABA [161], but diseased astrocytes around amyloid plaques become reactive and aberrantly and abundantly produce and release GABA via the anion channel bestrophin 1 (BEST1) [161] (Figure 3). Behaviorally, GABA from reactive astrocytes inhibits the activity of dentate granule neurons, resulting in inhibition of spike probability and contributing to learning and memory impairments [161].
A large body of literature indicates that AD disproportionally affects women in both occurrence and severity [162,163]. Interestingly, in a transgenic mouse model of AD (Tg2576 mice), GABA levels in the hippocampus were substantially higher in females than males [164]. This extra GABA was produced via the monoamine oxidase-B from putrescine in reactive astrocytes [164]. Thus, the reactivity of astrocytes and their dysregulated release of GABA may contribute to the observed sex differences in AD.

9. Altered D-Serine Release from Astrocytes in AD

NMDA receptors (NMDARs) play a central role in synaptic plasticity and learning, and memory. A distinctive feature of these receptors is the requirement for a co-agonist binding to the GluN1 subunit with glutamate binding to its recognition site on the GluN2-3 subunits for the activation of the receptor [165]. While glycine was originally thought to be the co-agonist on NMDARs, several studies over the last few decades have demonstrated that D-serine is at least as important in modulating NMDARs [166,167].
Serine racemase (SR) synthesizes D-serine in the brain by converting L-serine to D-serine. D-serine is a putative gliotransmitter that has been linked to learning and memory by its actions on synaptic NMDARs [168]. The astrocytic release of D-serine has been linked to the induction of LTP in the hippocampus and hypothalamus [168]. Altered levels of D-serine have been associated with neurological disorders, including schizophrenia and epilepsy. Interestingly, D-serine levels in postmortem hippocampal and cortical samples of AD patients seem to be higher than in control patients [169].
Moreover, levels of both D-serine and serine racemase, the enzyme responsible, are also elevated in experimental models of AD [169,170]. Increased expression of serine racemase in reactive astrocytes in TgF344-AD rats was also associated with elevated extra-synaptic NMDAR signaling [170]. The increased levels of D-serine and NMDAR hyperfunction may contribute to memory impairments and excitotoxicity observed in AD. More recently, it has been reported that serum D-serine levels serve as a biomarker for the progression of AD [171].

10. Changes to Gliotransmission in Neurodegenerative Diseases Other Than AD

It is not unexpected that altered astrocyte–neuronal communication and gliotransmission are also found in neurodegenerative diseases other than AD. Indeed, impaired glutamate uptake by astrocytes is a well-documented dysfunction of astrocytes in Parkinson’s disease (PD), amyotrophic lateral sclerosis (ALS), and Huntington’s disease (HD) [172]. Diminished levels of GLT-1 expression are found in postmortem brain tissues from patients with ALS, PD, or HD [173,174]. As discussed above, dysregulation of expression and function of GLT-1 is linked to chronically high levels of extracellular glutamate causing excitotoxicity [174]. Therefore, pharmacological induced upregulation of GLT-1 in neurodegenerative diseases is a promising approach to treating chronic excitotoxicity. Likewise, the enhanced expression of astrocytic Cx43 is altered in animal models of ALS and PD [175]. As in AD, the high expression of Cx43 is hypothesized to cause excessive release of ATP or glutamate into the extracellular space [175], ultimately leading or contributing to neurodegeneration [176]. Thus, there are many shared features in astrocyte–neuronal communication changes in diverse types of neurodegenerative diseases that may be a consequence of the astrocytic responses to their reactive state. Finding the commonalities and differences in astrocyte–neuronal communication in various neurodegenerative diseases may allow the design of novel strategies to treat these diseases.

11. Concluding Remarks

The exquisite and complex electrochemical communication between astrocytes and neurons is now being deciphered by the ever more sophisticated imaging and molecular techniques. It has become increasingly evident that this bidirectional communication between glia and neurons has striking effects on the functioning of synapses. However, changes in this form of communication during disease have received little consideration. We would argue from the above-mentioned studies that considerable evidence has now been gathered from many laboratories that imply that this glia–neuron dialogue is altered in AD with consequences for neuronal network functions (Figure 4). This impaired communication at synapses may well turn out to be an important factor in the chain of events that lead to cognitive impairments observed in AD. From a translational point of view, astrocytes may then provide important targets for therapeutic strategies for the treatment of AD.

Author Contributions

C.N. and A.M.B. wrote the initial draft of the manuscript, P.K. and A.A. edited and finalized the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Institutes of Health-NINDS (R01NS097312), National Institutes of Health-NIDA (R01DA048822), National Institutes of Health-NIMH (R01MH119355), and Postdoctoral Research Fellowship from the Basque Government (POS_2019_0041).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Dana Deters for expert laboratory assistance. Figures created in Biorender, https://biorender.com, 9 June 2021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hersi, M.; Irvine, B.; Gupta, P.; Gomes, J.; Birkett, N.; Krewski, D. Risk factors associated with the onset and progression of Alzheimer’s disease: A systematic review of the evidence. Neurotoxicology 2017, 61, 143–187. [Google Scholar] [CrossRef]
  2. Herrup, K. Reimagining Alzheimer’s disease—An age-based hypothesis. J. Neurosci. 2010, 30, 16755–16762. [Google Scholar] [CrossRef] [Green Version]
  3. Armstrong, R. Risk factors for Alzheimer’s disease. Folia Neuropathol. 2019, 57, 87–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hardy, J.; Selkoe, D.J. The amyloid hypothesis of Alzheimer’s disease: Progress and problems on the road to therapeutics. Science 2002, 297, 353–356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Stelzma, R.A.; Schnitzlein, H.N.; Murllagh, F.R. An English translation of Alzheimer’s 1907 paper, “Uber eine eigenartige Erkankung der Hirnrinde”. Clin. Anat. 1995, 8, 429–431. [Google Scholar] [CrossRef] [PubMed]
  6. Kumar, A.; Singh, A.; Ekavali, A. A review on Alzheimer’s disease pathophysiology and its management: An update. Pharmacol. Rep. 2015, 67, 195–203. [Google Scholar] [CrossRef] [PubMed]
  7. Francis, P.T.; Palmer, A.M.; Snape, M.; Wilcock, G.K. The cholinergic hypothesis of Alzheimer’s disease: A review of progress. J. Neurol. Neurosurg. Psychiatry 1999, 66, 137–147. [Google Scholar] [CrossRef]
  8. Drachman, D.A.; Leavitt, J. Human Memory and the Cholinergic System: A Relationship to Aging? Arch. Neurol. 1974, 30, 113–121. [Google Scholar] [CrossRef]
  9. Hardy, J.; Allsop, D. Amyloid deposition as the central event in the aetiology of Alzheimer’s disease. Trends Pharm. Sci. 1991, 12, 383–388. [Google Scholar] [CrossRef]
  10. Tanzi, R.E.; Bertram, L. Twenty Years of the Alzheimer’s Disease Amyloid Hypothesis: A Genetic Perspective. Cell 2005, 120, 545–555. [Google Scholar] [CrossRef]
  11. Hall, A.M.; Roberson, E.D. Mouse models of Alzheimer’s disease. Brain Res. Bull. 2012, 88, 3–12. [Google Scholar] [CrossRef] [Green Version]
  12. Small, S.A.; Duff, K. Linking Aβ and Tau in Late-Onset Alzheimer’s Disease: A Dual Pathway Hypothesis. Neuron 2008, 60, 534–542. [Google Scholar] [CrossRef] [Green Version]
  13. Kametani, F.; Hasegawa, M. Reconsideration of Amyloid Hypothesis and Tau Hypothesis in Alzheimer’s Disease. Front. Neurosci. 2018, 12, 25. [Google Scholar] [CrossRef] [Green Version]
  14. Kelley, B.J.; Petersen, R.C. Alzheimer’s disease and mild cognitive impairment. Neurol. Clin. 2007, 25, 577–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Fakhoury, M. Microglia and Astrocytes in Alzheimer’s Disease: Implications for Therapy. Curr. Neuropharmacol. 2018, 16, 508–518. [Google Scholar] [CrossRef] [PubMed]
  16. Phillips, E.C.; Croft, C.L.; Kurbatskaya, K.; O’Neill, M.J.; Hutton, M.L.; Hanger, D.P.; Garwood, C.J.; Noble, W. Astrocytes and neuroinflammation in Alzheimer’s disease. Biochem. Soc. Trans. 2014, 42, 1321–1325. [Google Scholar] [CrossRef] [PubMed]
  17. Ardura-Fabregat, A.; Boddeke, E.W.G.M.; Boza-Serrano, A.; Brioschi, S.; Castro-Gomez, S.; Ceyzériat, K.; Dansokho, C.; Dierkes, T.; Gelders, G.; Heneka, M.T.; et al. Targeting Neuroinflammation to Treat Alzheimer’s Disease. CNS Drugs 2017, 31, 1057–1082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Thal, D.R. The role of astrocytes in amyloid β-protein toxicity and clearance. Exp. Neurol. 2012, 236, 1–5. [Google Scholar] [CrossRef]
  19. Tarantini, S.; Tran, C.H.T.; Gordon, G.R.; Ungvari, Z.; Csiszar, A. Impaired neurovascular coupling in aging and Alzheimer’s disease: Contribution of astrocyte dysfunction and endothelial impairment to cognitive decline. Exp. Gerontol. 2017, 94, 52–58. [Google Scholar] [CrossRef]
  20. Kisler, K.; Nelson, A.R.; Montagne, A.; Zlokovic, B.V. Cerebral blood flow regulation and neurovascular dysfunction in Alzheimer disease. Nat. Rev. Neurosci. 2017, 18, 419. [Google Scholar] [CrossRef] [Green Version]
  21. Parpura, V.; Verkhratsky, A. Astrocytes revisited: Concise historic outlook on glutamate homeostasis and signaling. Croat. Med. J. 2012, 53, 518–528. [Google Scholar] [CrossRef] [Green Version]
  22. Von Bartheld, C.S.; Bahney, J.; Herculano-Houzel, S. The search for true numbers of neurons and glial cells in the human brain: A review of 150 years of cell counting. J. Comp. Neurol. 2016, 524, 3865–3895. [Google Scholar] [CrossRef] [Green Version]
  23. Miller, R.; Raff, M. Fibrous and protoplasmic astrocytes are biochemically and developmentally distinct. J. Neurosci. 1984, 4, 585–592. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, Y.; Barres, B.A. Astrocyte heterogeneity: An underappreciated topic in neurobiology. Curr. Opin. Neurobiol. 2010, 20, 588–594. [Google Scholar] [CrossRef] [PubMed]
  25. Xin, W.; Bonci, A. Functional Astrocyte Heterogeneity and Implications for Their Role in Shaping Neurotransmission. Front. Cell. Neurosci. 2018, 12, 141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Westergard, T.; Rothstein, J.D. Astrocyte Diversity: Current Insights and Future Directions. Neurochem. Res. 2020, 45, 1298–1305. [Google Scholar] [CrossRef]
  27. Bushong, E.A.; Martone, M.E.; Jones, Y.Z.; Ellisman, M.H. Protoplasmic astrocytes in CA1 stratum radiatum occupy separate anatomical domains. J. Neurosci. 2002, 22, 183–192. [Google Scholar] [CrossRef]
  28. Ogata, K.; Kosaka, T. Structural and quantitative analysis of astrocytes in the mouse hippocampus. Neuroscience 2002, 113, 221–233. [Google Scholar] [CrossRef]
  29. Halassa, M.M.; Fellin, T.; Takano, H.; Dong, J.-H.; Haydon, P.G. Synaptic islands defined by the territory of a single astrocyte. J. Neurosci. 2007, 27, 6473–6477. [Google Scholar] [CrossRef] [Green Version]
  30. Oberheim, N.A.; Takano, T.; Han, X.; He, W.; Lin, J.H.C.; Wang, F.; Xu, Q.; Wyatt, J.D.; Pilcher, W.; Ojemann, J.G.; et al. Uniquely Hominid Features of Adult Human Astrocytes. J. Neurosci. 2009, 29, 3276–3287. [Google Scholar] [CrossRef] [PubMed]
  31. Anderson, C.M.; Swanson, R.A. Astrocyte glutamate transport: Review of properties, regulation, and physiological functions. Glia 2000, 32, 1–14. [Google Scholar] [CrossRef]
  32. Schousboe, A. Pharmacological and Functional Characterization of Astrocytic GABA Transport: A Short Review. Neurochem. Res. 2000, 25, 1241–1244. [Google Scholar] [CrossRef]
  33. Sofroniew, M.V.; Vinters, H.V. Astrocytes: Biology and pathology. Acta Neuropathol. 2010, 119, 7–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Magaki, S.D.; Williams, C.K.; Vinters, H.V. Glial function (and dysfunction) in the normal & ischemic brain. Neuropharmacology 2018, 134, 218–225. [Google Scholar]
  35. Farhy-Tselnicker, I.; Allen, N.J. Astrocytes, neurons, synapses: A tripartite view on cortical circuit development. Neural Dev. 2018, 13, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Clarke, L.E.; Barres, B.A. Emerging roles of astrocytes in neural circuit development. Nat. Rev. Neurosci. 2013, 14, 311–321. [Google Scholar] [CrossRef] [Green Version]
  37. Khakh, B.S.; Sofroniew, M.V. Diversity of astrocyte functions and phenotypes in neural circuits. Nat. Neurosci. 2015, 18, 942–952. [Google Scholar] [CrossRef]
  38. Kofuji, P.; Newman, E.A. Potassium buffering in the central nervous system. Neuroscience 2004, 129, 1045–1056. [Google Scholar] [CrossRef] [Green Version]
  39. Nwaobi, S.E.; Cuddapah, V.A.; Patterson, K.C.; Randolph, A.C.; Olsen, M.L. The role of glial-specific Kir4.1 in normal and pathological states of the CNS. Acta Neuropathol. 2016, 132, 1–21. [Google Scholar] [CrossRef]
  40. Olsen, M.L.; Sontheimer, H. Functional implications for Kir4.1 channels in glial biology: From K+ buffering to cell differentiation. J. Neurochem. 2008, 107, 589–601. [Google Scholar] [CrossRef] [Green Version]
  41. Nagy, J.I.; Rash, J.E. Connexins and gap junctions of astrocytes and oligodendrocytes in the CNS. Brain Res. Rev. 2000, 32, 29–44. [Google Scholar] [CrossRef]
  42. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef] [PubMed]
  43. Santello, M.; Calì, C.; Bezzi, P. Gliotransmission and the Tripartite Synapse. In Synaptic Plasticity: Dynamics, Development and Disease; Kreutz, M.R., Sala, C., Eds.; Springer: Vienna, Austria, 2012; pp. 307–331. [Google Scholar]
  44. Araque, A.; Parpura, V.; Sanzgiri, R.P.; Haydon, P.G. Tripartite synapses: Glia, the unacknowledged partner. Trends Neurosci. 1999, 22, 208–215. [Google Scholar] [CrossRef]
  45. Durkee, C.A.; Araque, A. Diversity and Specificity of Astrocyte–neuron Communication. Neuroscience 2019, 396, 73–78. [Google Scholar] [CrossRef]
  46. Kofuji, P.; Araque, A. G-Protein-Coupled Receptors in Astrocyte–Neuron Communication. Neuroscience 2020, 456, 71–84. [Google Scholar] [CrossRef] [PubMed]
  47. Araque, A.; Carmignoto, G.; Haydon, P.G.; Oliet, S.H.; Robitaille, R.; Volterra, A. Gliotransmitters travel in time and space. Neuron 2014, 81, 728–739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Bazargani, N.; Attwell, D. Astrocyte calcium signaling: The third wave. Nat. Neurosci. 2016, 19, 182–189. [Google Scholar] [CrossRef] [PubMed]
  49. Lines, J.; Martin, E.D.; Kofuji, P.; Aguilar, J.; Araque, A. Astrocytes modulate sensory-evoked neuronal network activity. Nat. Commun. 2020, 11, 3689. [Google Scholar] [CrossRef]
  50. Volterra, A.; Meldolesi, J. Astrocytes, from brain glue to communication elements: The revolution continues. Nat. Rev. Neurosci. 2005, 6, 626–640. [Google Scholar] [CrossRef]
  51. Perea, G.; Araque, A. Astrocytes potentiate transmitter release at single hippocampal synapses. Science 2007, 317, 1083–1086. [Google Scholar] [CrossRef]
  52. Navarrete, M.; Araque, A. Endocannabinoids potentiate synaptic transmission through stimulation of astrocytes. Neuron 2010, 68, 113–126. [Google Scholar] [CrossRef] [Green Version]
  53. Parri, H.R.; Gould, T.M.; Crunelli, V. Spontaneous astrocytic Ca2+ oscillations in situ drive NMDAR-mediated neuronal excitation. Nat. Neurosci. 2001, 4, 803–812. [Google Scholar] [CrossRef]
  54. Bezzi, P.; Gundersen, V.; Galbete, J.L.; Seifert, G.; Steinhauser, C.; Pilati, E.; Volterra, A. Astrocytes contain a vesicular compartment that is competent for regulated exocytosis of glutamate. Nat. Neurosci. 2004, 7, 613–620. [Google Scholar] [CrossRef]
  55. Martin, R.; Bajo-Graneras, R.; Moratalla, R.; Perea, G.; Araque, A. Circuit-specific signaling in astrocyte-neuron networks in basal ganglia pathways. Science 2015, 349, 730–734. [Google Scholar] [CrossRef] [Green Version]
  56. Panatier, A.; Vallee, J.; Haber, M.; Murai, K.K.; Lacaille, J.C.; Robitaille, R. Astrocytes are endogenous regulators of basal transmission at central synapses. Cell 2011, 146, 785–798. [Google Scholar] [CrossRef] [Green Version]
  57. Di Castro, M.A.; Chuquet, J.; Liaudet, N.; Bhaukaurally, K.; Santello, M.; Bouvier, D.; Tiret, P.; Volterra, A. Local Ca2+ detection and modulation of synaptic release by astrocytes. Nat. Neurosci. 2011, 14, 1276–1284. [Google Scholar] [CrossRef] [PubMed]
  58. Min, R.; Nevian, T. Astrocyte signaling controls spike timing-dependent depression at neocortical synapses. Nat. Neurosci. 2012, 15, 746–753. [Google Scholar] [CrossRef] [PubMed]
  59. Halassa, M.M.; Haydon, P.G. Integrated brain circuits: Astrocytic networks modulate neuronal activity and behavior. Annu. Rev. Physiol. 2010, 72, 335–355. [Google Scholar] [CrossRef] [Green Version]
  60. Perea, G.; Gomez, R.; Mederos, S.; Covelo, A.; Ballesteros, J.J.; Schlosser, L.; Hernandez-Vivanco, A.; Martin-Fernandez, M.; Quintana, R.; Rayan, A.; et al. Activity-dependent switch of GABAergic inhibition into glutamatergic excitation in astrocyte-neuron networks. Elife 2016, 5, e20362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Poskanzer, K.E.; Yuste, R. Astrocytes regulate cortical state switching in vivo. Proc. Natl. Acad. Sci. USA 2016, 113, E2675–E2684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lee, H.S.; Ghetti, A.; Pinto-Duarte, A.; Wang, X.; Dziewczapolski, G.; Galimi, F.; Huitron-Resendiz, S.; Pina-Crespo, J.C.; Roberts, A.J.; Verma, I.M.; et al. Astrocytes contribute to gamma oscillations and recognition memory. Proc. Natl. Acad. Sci. USA 2014, 111, E3343–E3352. [Google Scholar] [CrossRef] [Green Version]
  63. Oliveira, J.F.; Sardinha, V.M.; Guerra-Gomes, S.; Araque, A.; Sousa, N. Do stars govern our actions? Astrocyte involvement in rodent behavior. Trends Neurosci. 2015, 38, 535–549. [Google Scholar] [CrossRef] [Green Version]
  64. Xie, A.X.; Petravicz, J.; McCarthy, K.D. Molecular approaches for manipulating astrocytic signaling in vivo. Front. Cell. Neurosci. 2015, 9, 144. [Google Scholar] [CrossRef] [Green Version]
  65. Kofuji, P.; Araque, A. Astrocytes and Behavior. Annu. Rev. Neurosci. 2021, 44, 49–67. [Google Scholar] [CrossRef]
  66. De Pittà, M.; Brunel, N.; Volterra, A. Astrocytes: Orchestrating synaptic plasticity? Neuroscience 2016, 323, 43–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Sofroniew, M.V. Astrocyte Reactivity: Subtypes, States, and Functions in CNS Innate Immunity. Trends Immunol. 2020, 41, 758–770. [Google Scholar] [CrossRef] [PubMed]
  68. Li, K.; Li, J.; Zheng, J.; Qin, S. Reactive Astrocytes in Neurodegenerative Diseases. Aging Dis. 2019, 10, 664–675. [Google Scholar] [CrossRef] [Green Version]
  69. Matias, I.; Morgado, J.; Gomes, F.C.A. Astrocyte Heterogeneity: Impact to Brain Aging and Disease. Front. Aging Neurosci. 2019, 11, 59. [Google Scholar] [CrossRef] [Green Version]
  70. Eng, L.F.; Ghirnikar, R.S. GFAP and Astrogliosis. Brain Pathol. 1994, 4, 229–237. [Google Scholar] [CrossRef]
  71. Pekny, M.; Nilsson, M. Astrocyte activation and reactive gliosis. Glia 2005, 50, 427–434. [Google Scholar] [CrossRef]
  72. Chun, H.; Lee, C.J. Reactive astrocytes in Alzheimer’s disease: A double-edged sword. Neurosci. Res. 2018, 126, 44–52. [Google Scholar] [CrossRef] [PubMed]
  73. Pike, C.J.; Cummings, B.J.; Cotman, C.W. Early association of reactive astrocytes with senile plaques in Alzheimer’s disease. Exp. Neurol. 1995, 132, 172–179. [Google Scholar] [CrossRef]
  74. Perez-Nievas, B.G.; Serrano-Pozo, A. Deciphering the Astrocyte Reaction in Alzheimer’s Disease. Front. Aging Neurosci. 2018, 10, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Leng, K.; Li, E.; Eser, R.; Piergies, A.; Sit, R.; Tan, M.; Neff, N.; Li, S.H.; Rodriguez, R.D.; Suemoto, C.K.; et al. Molecular characterization of selectively vulnerable neurons in Alzheimer’s disease. Nat. Neurosci. 2021, 24, 276–287. [Google Scholar] [CrossRef]
  76. Agulhon, C.; Sun, M.-Y.; Murphy, T.; Myers, T.; Lauderdale, K.; Fiacco, T. Calcium Signaling and Gliotransmission in Normal vs. Reactive Astrocytes. Front. Pharmacol. 2012, 3, 139. [Google Scholar] [CrossRef] [Green Version]
  77. Delekate, A.; Füchtemeier, M.; Schumacher, T.; Ulbrich, C.; Foddis, M.; Petzold, G.C. Metabotropic P2Y1 receptor signalling mediates astrocytic hyperactivity in vivo in an Alzheimer’s disease mouse model. Nat. Commun. 2014, 5, 5422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kuchibhotla, K.V.; Lattarulo, C.R.; Hyman, B.T.; Bacskai, B.J. Synchronous hyperactivity and intercellular calcium waves in astrocytes in Alzheimer mice. Science 2009, 323, 1211–1215. [Google Scholar] [CrossRef] [Green Version]
  79. Pirttimaki, T.M.; Codadu, N.K.; Awni, A.; Pratik, P.; Nagel, D.A.; Hill, E.J.; Dineley, K.T.; Parri, H.R. α7 Nicotinic Receptor-Mediated Astrocytic Gliotransmitter Release: Aβ Effects in a Preclinical Alzheimer’s Mouse Model. PLoS ONE 2013, 8, e81828. [Google Scholar] [CrossRef]
  80. Reichenbach, N.; Delekate, A.; Breithausen, B.; Keppler, K.; Poll, S.; Schulte, T.; Peter, J.; Plescher, M.; Hansen, J.N.; Blank, N.; et al. P2Y1 receptor blockade normalizes network dysfunction and cognition in an Alzheimer’s disease model. J. Exp. Med. 2018, 215, 1649–1663. [Google Scholar] [CrossRef] [PubMed]
  81. Takano, T.; Han, X.; Deane, R.; Zlokovic, B.; Nedergaard, M. Two-photon imaging of astrocytic Ca2+ signaling and the microvasculature in experimental mice models of Alzheimer’s disease. Ann. N. Y. Acad. Sci. 2007, 1097, 40–50. [Google Scholar] [CrossRef] [PubMed]
  82. Yi, C.; Mei, X.; Ezan, P.; Mato, S.; Matias, I.; Giaume, C.; Koulakoff, A. Astroglial connexin43 contributes to neuronal suffering in a mouse model of Alzheimer’s disease. Cell Death Differ. 2016, 23, 1691–1701. [Google Scholar] [CrossRef]
  83. Gómez-Gonzalo, M.; Martin-Fernandez, M.; Martínez-Murillo, R.; Mederos, S.; Hernández-Vivanco, A.; Jamison, S.; Fernandez, A.P.; Serrano, J.; Calero, P.; Futch, H.S.; et al. Neuron-astrocyte signaling is preserved in the aging brain. Glia 2017, 65, 569–580. [Google Scholar] [CrossRef] [Green Version]
  84. Lalo, U.; Bogdanov, A.; Pankratov, Y. Diversity of Astroglial Effects on Aging- and Experience-Related Cortical Metaplasticity. Front. Mol. Neurosci. 2018, 11, 239. [Google Scholar] [CrossRef] [PubMed]
  85. Erb, L.; Cao, C.; Ajit, D.; Weisman, G.A. P2Y receptors in Alzheimer’s disease. Biol. Cell 2015, 107, 1–21. [Google Scholar] [CrossRef] [Green Version]
  86. Teaktong, T.; Graham, A.; Court, J.; Perry, R.; Jaros, E.; Johnson, M.; Hall, R.; Perry, E. Alzheimer’s disease is associated with a selective increase in α7 nicotinic acetylcholine receptor immunoreactivity in astrocytes. Glia 2003, 41, 207–211. [Google Scholar] [CrossRef]
  87. Yu, W.-F.; Guan, Z.-Z.; Bogdanovic, N.; Nordberg, A. High selective expression of α7 nicotinic receptors on astrocytes in the brains of patients with sporadic Alzheimer’s disease and patients carrying Swedish APP 670/671 mutation: A possible association with neuritic plaques. Exp. Neurol. 2005, 192, 215–225. [Google Scholar] [CrossRef]
  88. Pham, C.; Hérault, K.; Oheim, M.; Maldera, S.; Vialou, V.; Cauli, B.; Li, D. Astrocytes respond to a neurotoxic Aβ fragment with state-dependent Ca2+ alteration and multiphasic transmitter release. Acta Neuropathol. Commun. 2021, 9, 44. [Google Scholar] [CrossRef] [PubMed]
  89. Parri, H.R.; Hernandez, C.M.; Dineley, K.T. Research update: Alpha7 nicotinic acetylcholine receptor mechanisms in Alzheimer’s disease. Biochem. Pharmacol. 2011, 82, 931–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Lee, L.; Kosuri, P.; Arancio, O. Picomolar amyloid-β peptides enhance spontaneous astrocyte calcium transients. J. Alzheimers Dis. 2014, 38, 49–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Fucile, S. Ca2+ permeability of nicotinic acetylcholine receptors. Cell Calcium 2004, 35, 1–8. [Google Scholar] [CrossRef] [PubMed]
  92. Sharma, G.; Vijayaraghavan, S. Nicotinic cholinergic signaling in hippocampal astrocytes involves calcium-induced calcium release from intracellular stores. Proc. Natl. Acad. Sci. USA 2001, 98, 4148–4153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Panatier, A.; Robitaille, R. Astrocytic mGluR5 and the tripartite synapse. Neuroscience 2016, 323, 29–34. [Google Scholar] [CrossRef] [PubMed]
  94. Sun, W.; McConnell, E.; Pare, J.-F.; Xu, Q.; Chen, M.; Peng, W.; Lovatt, D.; Han, X.; Smith, Y.; Nedergaard, M. Glutamate-Dependent Neuroglial Calcium Signaling Differs Between Young and Adult Brain. Science 2013, 339, 197–200. [Google Scholar] [CrossRef] [Green Version]
  95. Aronica, E.; Van Vliet, E.A.; Mayboroda, O.A.; Troost, D.; Da Silva, F.H.L.; Gorter, J.A. Upregulation of metabotropic glutamate receptor subtype mGluR3 and mGluR5 in reactive astrocytes in a rat model of mesial temporal lobe epilepsy. Eur. J. Neurosci. 2000, 12, 2333–2344. [Google Scholar] [CrossRef] [PubMed]
  96. Aronica, E.; Catania, M.V.; Geurts, J.; Yankaya, B.; Troost, D. Immunohistochemical localization of group I and II metabotropic glutamate receptors in control and amyotrophic lateral sclerosis human spinal cord: Upregulation in reactive astrocytes. Neuroscience 2001, 105, 509–520. [Google Scholar] [CrossRef]
  97. Planas-Fontánez, T.M.; Dreyfus, C.F.; Saitta, K.S. Reactive Astrocytes as Therapeutic Targets for Brain Degenerative Diseases: Roles Played by Metabotropic Glutamate Receptors. Neurochem. Res. 2020, 45, 541–550. [Google Scholar] [CrossRef] [Green Version]
  98. Casley, C.S.; Lakics, V.; Lee, H.-G.; Broad, L.M.; Day, T.A.; Cluett, T.; Smith, M.A.; O’Neill, M.J.; Kingston, A.E. Up-regulation of astrocyte metabotropic glutamate receptor 5 by amyloid-β peptide. Brain Res. 2009, 1260, 65–75. [Google Scholar] [CrossRef]
  99. Lim, D.; Iyer, A.; Ronco, V.; Grolla, A.A.; Canonico, P.L.; Aronica, E.; Genazzani, A.A. Amyloid beta deregulates astroglial mGluR5-mediated calcium signaling via calcineurin and Nf-kB. Glia 2013, 61, 1134–1145. [Google Scholar] [CrossRef]
  100. Wevers, A.; Schröder, H. Nicotinic Acetylcholine Receptors in Alzheimer’s Disease. J. Alzheimers Dis. 1999, 1, 207–219. [Google Scholar] [CrossRef] [PubMed]
  101. Rodríguez-Arellano, J.J.; Parpura, V.; Zorec, R.; Verkhratsky, A. Astrocytes in physiological aging and Alzheimer’s disease. Neuroscience 2016, 323, 170–182. [Google Scholar] [CrossRef] [PubMed]
  102. Dossi, E.; Vasile, F.; Rouach, N. Human astrocytes in the diseased brain. Brain Res. Bull. 2018, 136, 139–156. [Google Scholar] [CrossRef]
  103. Grolla, A.A.; Fakhfouri, G.; Balzaretti, G.; Marcello, E.; Gardoni, F.; Canonico, P.L.; Di Luca, M.; Genazzani, A.A.; Lim, D. Aβ leads to Ca2+ signaling alterations and transcriptional changes in glial cells. Neurobiol. Aging 2013, 34, 511–522. [Google Scholar] [CrossRef] [PubMed]
  104. 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]
  105. Busche, M.A.; Eichhoff, G.; Adelsberger, H.; Abramowski, D.; Wiederhold, K.-H.; Haass, C.; Staufenbiel, M.; Konnerth, A.; Garaschuk, O. Clusters of Hyperactive Neurons Near Amyloid Plaques in a Mouse Model of Alzheimer’s Disease. Science 2008, 321, 1686–1689. [Google Scholar] [CrossRef] [Green Version]
  106. Kuchibhotla, K.V.; Goldman, S.T.; Lattarulo, C.R.; Wu, H.-Y.; Hyman, B.T.; Bacskai, B.J. Aβ Plaques Lead to Aberrant Regulation of Calcium Homeostasis In Vivo Resulting in Structural and Functional Disruption of Neuronal Networks. Neuron 2008, 59, 214–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Richetin, K.; Steullet, P.; Pachoud, M.; Perbet, R.; Parietti, E.; Maheswaran, M.; Eddarkaoui, S.; Bégard, 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] [PubMed]
  108. Reiner, A.; Levitz, J. Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert. Neuron 2018, 98, 1080–1098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Murphy-Royal, C.; Dupuis, J.; Groc, L.; Oliet, S.H.R. Astroglial glutamate transporters in the brain: Regulating neurotransmitter homeostasis and synaptic transmission. J. Neurosci. Res. 2017, 95, 2140–2151. [Google Scholar] [CrossRef] [Green Version]
  110. Rothstein, J.D.; Dykes-Hoberg, M.; Pardo, C.A.; Bristol, L.A.; Jin, L.; Kuncl, R.W.; Kanai, Y.; Hediger, M.A.; Wang, Y.; Schielke, J.P.; et al. Knockout of Glutamate Transporters Reveals a Major Role for Astroglial Transport in Excitotoxicity and Clearance of Glutamate. Neuron 1996, 16, 675–686. [Google Scholar] [CrossRef] [Green Version]
  111. Lehre, K.P.; Danbolt, N.C. The Number of Glutamate Transporter Subtype Molecules at Glutamatergic Synapses: Chemical and Stereological Quantification in Young Adult Rat Brain. J. Neurosci. 1998, 18, 8751–8757. [Google Scholar] [CrossRef] [Green Version]
  112. Tanaka, K.; Watase, K.; Manabe, T.; Yamada, K.; Watanabe, M.; Takahashi, K.; Iwama, H.; Nishikawa, T.; Ichihara, N.; Kikuchi, T.; et al. Epilepsy and exacerbation of brain injury in mice lacking the glutamate transporter GLT-1. Science 1997, 276, 1699–1702. [Google Scholar] [CrossRef] [PubMed]
  113. Acioglu, C.; Li, L.; Elkabes, S. Contribution of astrocytes to neuropathology of neurodegenerative diseases. Brain Res. 2021, 1758, 147291. [Google Scholar] [CrossRef]
  114. Beckstrøm, H.; Julsrud, L.; Haugeto, O.; Dewar, D.; Graham, D.I.; Lehre, K.P.; Storm-Mathisen, J.; Danbolt, N.C. Interindividual differences in the levels of the glutamate transporters GLAST and GLT, but no clear correlation with Alzheimer’s disease. J. Neurosci. Res. 1999, 55, 218–229. [Google Scholar] [CrossRef]
  115. Garcia-Esparcia, P.; Diaz-Lucena, D.; Ainciburu, M.; Torrejón-Escribano, B.; Carmona, M.; Llorens, F.; Ferrer, I. Glutamate Transporter GLT1 Expression in Alzheimer Disease and Dementia With Lewy Bodies. Front. Aging Neurosci. 2018, 10, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Peters, O.; Schipke, C.G.; Philipps, A.; Haas, B.; Pannasch, U.; Wang, L.P.; Benedetti, B.; Kingston, A.E.; Kettenmann, H. Astrocyte function is modified by alzheimer’s disease-like pathology in aged mice. J. Alzheimers Dis. 2009, 18, 177–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Masliah, E.; Alford, M.; Mallory, M.; Rockenstein, E.; Moechars, D.; Van Leuven, F. Abnormal Glutamate Transport Function in Mutant Amyloid Precursor Protein Transgenic Mice. Exp. Neurol. 2000, 163, 381–387. [Google Scholar] [CrossRef] [PubMed]
  118. Li, S.; Mallory, M.; Alford, M.; Tanaka, S.; Masliah, E. Glutamate transporter alterations in Alzheimer disease are possibly associated with abnormal APP expression. J. Neuropathol. Exp. Neurol. 1997, 56, 901–911. [Google Scholar] [CrossRef]
  119. Jacob, C.P.; Koutsilieri, E.; Bartl, J.; Neuen-Jacob, E.; Arzberger, T.; Zander, N.; Ravid, R.; Roggendorf, W.; Riederer, P.; Grünblatt, E. Alterations in Expression of Glutamatergic Transporters and Receptors in Sporadic Alzheimer’s Disease. J. Alzheimers Dis. 2007, 11, 97–116. [Google Scholar] [CrossRef]
  120. Masliah, E.; Alford, M.; DeTeresa, R.; Mallory, M.; Hansen, L. Deficient glutamate transport is associated with neurodegeneration in Alzheimer’s disease. Ann. Neurol. 1996, 40, 759–766. [Google Scholar] [CrossRef] [PubMed]
  121. Mookherjee, P.; Green, P.S.; Watson, G.S.; Marques, M.A.; Tanaka, K.; Meeker, K.D.; Meabon, J.S.; Li, N.; Zhu, P.; Olson, V.G.; et al. GLT-1 loss accelerates cognitive deficit onset in an Alzheimer’s disease animal model. J. Alzheimers Dis. 2011, 26, 447–455. [Google Scholar] [CrossRef]
  122. Takahashi, K.; Kong, Q.; Lin, Y.; Stouffer, N.; Schulte, D.A.; Lai, L.; Liu, Q.; Chang, L.-C.; Dominguez, S.; Xing, X.; et al. Restored glial glutamate transporter EAAT2 function as a potential therapeutic approach for Alzheimer’s disease. J. Exp. Med. 2015, 212, 319–332. [Google Scholar] [CrossRef]
  123. Hefendehl, J.K.; LeDue, J.; Ko, R.W.; Mahler, J.; Murphy, T.H.; MacVicar, B.A. Mapping synaptic glutamate transporter dysfunction in vivo to regions surrounding Aβ plaques by iGluSnFR two-photon imaging. Nat. Commun. 2016, 7, 13441. [Google Scholar] [CrossRef] [Green Version]
  124. Kobayashi, E.; Nakano, M.; Kubota, K.; Himuro, N.; Mizoguchi, S.; Chikenji, T.; Otani, M.; Mizue, Y.; Nagaishi, K.; Fujimiya, M. Activated forms of astrocytes with higher GLT-1 expression are associated with cognitive normal subjects with Alzheimer pathology in human brain. Sci. Rep. 2018, 8, 1712. [Google Scholar] [CrossRef]
  125. Parpura, V.; Zorec, R. Gliotransmission: Exocytotic release from astrocytes. Brain Res. Rev. 2010, 63, 83–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Santello, M.; Volterra, A. Synaptic modulation by astrocytes via Ca2+-dependent glutamate release. Neuroscience 2009, 158, 253–259. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, Q.-S.; Xu, Q.; Kang, J.; Nedergaard, M. Astrocyte activation of presynaptic metabotropic glutamate receptors modulates hippocampal inhibitory synaptic transmission. Neuron Glia Biol. 2004, 1, 307–316. [Google Scholar] [CrossRef] [PubMed]
  128. Angulo, E.; Noé, V.; Casadó, V.; Mallol, J.; Gomez-Isla, T.; Lluis, C.; Ferrer, I.; Ciudad, C.J.; Franco, R. Up-regulation of the Kv3.4 potassium channel subunit in early stages of Alzheimer’s disease. J. Neurochem. 2004, 91, 547–557. [Google Scholar] [CrossRef]
  129. Fellin, T.; Carmignoto, G. Neurone-to-astrocyte signalling in the brain represents a distinct multifunctional unit. J. Physiol. 2004, 559, 3–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Perea, G.; Araque, A. Properties of synaptically evoked astrocyte calcium signal reveal synaptic information processing by astrocytes. J. Neurosci. 2005, 25, 2192–2203. [Google Scholar] [CrossRef]
  131. Shigetomi, E.; Bowser, D.N.; Sofroniew, M.V.; Khakh, B.S. Two forms of astrocyte calcium excitability have distinct effects on NMDA receptor-mediated slow inward currents in pyramidal neurons. J. Neurosci. 2008, 28, 6659–6663. [Google Scholar] [CrossRef] [PubMed]
  132. Kovács, A.; Pál, B. Astrocyte-Dependent Slow Inward Currents (SICs) Participate in Neuromodulatory Mechanisms in the Pedunculopontine Nucleus (PPN). Front. Cell. Neurosci. 2017, 11, 16. [Google Scholar] [CrossRef] [Green Version]
  133. Angulo, M.C.; Kozlov, A.S.; Charpak, S.; Audinat, E. Glutamate released from glial cells synchronizes neuronal activity in the hippocampus. J. Neurosci. 2004, 24, 6920–6927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Fellin, T.; Pascual, O.; Gobbo, S.; Pozzan, T.; Haydon, P.G.; Carmignoto, G. Neuronal synchrony mediated by astrocytic glutamate through activation of extrasynaptic NMDA receptors. Neuron 2004, 43, 729–743. [Google Scholar] [CrossRef] [Green Version]
  135. Gajardo-Gómez, R.; Labra, V.C.; Maturana, C.J.; Shoji, K.F.; Santibañez, C.A.; Sáez, J.C.; Giaume, C.; Orellana, J.A. Cannabinoids prevent the amyloid β-induced activation of astroglial hemichannels: A neuroprotective mechanism. Glia 2017, 65, 122–137. [Google Scholar] [CrossRef] [PubMed]
  136. Orellana, J.A.; Froger, N.; Ezan, P.; Jiang, J.X.; Bennett, M.V.L.; Naus, C.C.; Giaume, C.; Sáez, J.C. ATP and glutamate released via astroglial connexin 43 hemichannels mediate neuronal death through activation of pannexin 1 hemichannels. J. Neurochem. 2011, 118, 826–840. [Google Scholar] [CrossRef] [Green Version]
  137. Talantova, M.; Sanz-Blasco, S.; Zhang, X.; Xia, P.; Akhtar, M.W.; Okamoto, S.-I.; Dziewczapolski, G.; Nakamura, T.; Cao, G.; Pratt, A.E.; et al. Aβ induces astrocytic glutamate release, extrasynaptic NMDA receptor activation, and synaptic loss. Proc. Natl. Acad. Sci. USA 2013, 110, E2518–E2527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Burnstock, G. Historical review: ATP as a neurotransmitter. Trends Pharmacol. Sci. 2006, 27, 166–176. [Google Scholar] [CrossRef]
  139. Harada, K.; Kamiya, T.; Tsuboi, T. Gliotransmitter release from astrocytes: Functional, developmental and pathological implications in the brain. Front. Neurosci. 2016, 9, 499. [Google Scholar] [CrossRef] [Green Version]
  140. Pangršič, T.; Potokar, M.; Stenovec, M.; Kreft, M.; Fabbretti, E.; Nistri, A.; Pryazhnikov, E.; Khiroug, L.; Giniatullin, R.; Zorec, R. Exocytotic Release of ATP from Cultured Astrocytes*. J. Biol. Chem. 2007, 282, 28749–28758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Lalo, U.; Palygin, O.; Rasooli-Nejad, S.; Andrew, J.; Haydon, P.G.; Pankratov, Y. Exocytosis of ATP From Astrocytes Modulates Phasic and Tonic Inhibition in the Neocortex. PLoS Biol. 2014, 12, e1001747. [Google Scholar] [CrossRef] [PubMed]
  142. Zhang, Z.; Chen, G.; Zhou, W.; Song, A.; Xu, T.; Luo, Q.; Wang, W.; Gu, X.-S.; Duan, S. Regulated ATP release from astrocytes through lysosome exocytosis. Nat. Cell Biol. 2007, 9, 945–953. [Google Scholar] [CrossRef]
  143. Liu, T.; Sun, L.; Xiong, Y.; Shang, S.; Guo, N.; Teng, S.; Wang, Y.; Liu, B.; Wang, C.; Wang, L.; et al. Calcium Triggers Exocytosis from Two Types of Organelles in a Single Astrocyte. J. Neurosci. 2011, 31, 10593–10601. [Google Scholar] [CrossRef] [Green Version]
  144. Verkhratsky, A.; Matteoli, M.; Parpura, V.; Mothet, J.-P.; Zorec, R. Astrocytes as secretory cells of the central nervous system: Idiosyncrasies of vesicular secretion. EMBO J. 2016, 35, 239–257. [Google Scholar] [CrossRef] [Green Version]
  145. Anselmi, F.; Hernandez, V.H.; Crispino, G.; Seydel, A.; Ortolano, S.; Roper, S.D.; Kessaris, N.; Richardson, W.; Rickheit, G.; Filippov, M.A.; et al. ATP release through connexin hemichannels and gap junction transfer of second messengers propagate Ca2+ signals across the inner ear. Proc. Natl. Acad. Sci. USA 2008, 105, 18770–18775. [Google Scholar] [CrossRef] [Green Version]
  146. Suadicani, S.O.; Iglesias, R.; Wang, J.; Dahl, G.; Spray, D.C.; Scemes, E. ATP signaling is deficient in cultured pannexin1-null mouse astrocytes. Glia 2012, 60, 1106–1116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Jun, M.; Xiaolong, Q.; Chaojuan, Y.; Ruiyuan, P.; Shukun, W.; Junbing, W.; Li, H.; Hong, C.; Jinbo, C.; Rong, W.; et al. Calhm2 governs astrocytic ATP releasing in the development of depression-like behaviors. Mol. Psychiatry 2018, 23, 883–891. [Google Scholar] [CrossRef]
  148. Haughey, N.J.; Mattson, M.P. Alzheimer’s amyloid β-peptide enhances ATP/gap junction-mediated calcium-wave propagation in astrocytes. Neuromol. Med. 2003, 3, 173–180. [Google Scholar] [CrossRef]
  149. Nagy, J.I.; Li, W.; Hertzberg, E.L.; Marotta, C.A. Elevated connexin43 immunoreactivity at sites of amyloid plaques in Alzheimer’s disease. Brain Res. 1996, 717, 173–178. [Google Scholar] [CrossRef]
  150. Angeli, S.; Kousiappa, I.; Stavrou, M.; Sargiannidou, I.; Georgiou, E.; Papacostas, S.S.; Kleopa, K.A. Altered Expression of Glial Gap Junction Proteins Cx43, Cx30, and Cx47 in the 5XFAD Model of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 1060. [Google Scholar] [CrossRef]
  151. Kajiwara, Y.; Wang, E.; Wang, M.; Sin, W.C.; Brennand, K.J.; Schadt, E.; Naus, C.C.; Buxbaum, J.; Zhang, B. GJA1 (connexin43) is a key regulator of Alzheimer’s disease pathogenesis. Acta Neuropathol. Commun. 2018, 6, 144. [Google Scholar] [CrossRef] [PubMed]
  152. Chen, W.T.; Lu, A.; Craessaerts, K.; Pavie, B.; Sala Frigerio, C.; Corthout, N.; Qian, X.; Laláková, J.; Kühnemund, M.; Voytyuk, I.; et al. Spatial Transcriptomics and In Situ Sequencing to Study Alzheimer’s Disease. Cell 2020, 182, 976–991. [Google Scholar] [CrossRef]
  153. Koulakoff, A.; Mei, X.; Orellana, J.A.; Sáez, J.C.; Giaume, C. Glial connexin expression and function in the context of Alzheimer’s disease. Biochim. Biophys. Acta 2012, 1818, 2048–2057. [Google Scholar] [CrossRef] [Green Version]
  154. Orellana, J.A.; Shoji, K.F.; Abudara, V.; Ezan, P.; Amigou, E.; Sáez, P.J.; Jiang, J.X.; Naus, C.C.; Sáez, J.C.; Giaume, C. Amyloid β-induced death in neurons involves glial and neuronal hemichannels. J. Neurosci. 2011, 31, 4962–4977. [Google Scholar] [CrossRef] [PubMed]
  155. Kang, J.; Kang, N.; Lovatt, D.; Torres, A.; Zhao, Z.; Lin, J.; Nedergaard, M. Connexin 43 hemichannels are permeable to ATP. J. Neurosci. 2008, 28, 4702–4711. [Google Scholar] [CrossRef]
  156. Jung, E.S.; An, K.; Seok Hong, H.; Kim, J.-H.; Mook-Jung, I. Astrocyte-Originated ATP Protects Aβ1-42-Induced Impairment of Synaptic Plasticity. J. Neurosci. 2012, 32, 3081–3087. [Google Scholar] [CrossRef] [Green Version]
  157. Yoon, B.-E.; Lee, C.J. GABA as a rising gliotransmitter. Front. Neural Circuits 2014, 8, 141. [Google Scholar] [CrossRef] [Green Version]
  158. Garaschuk, O.; Verkhratsky, A. GABAergic astrocytes in Alzheimer’s disease. Aging 2019, 11, 1602–1604. [Google Scholar] [CrossRef] [PubMed]
  159. Yoon, B.E.; Woo, J.; Chun, Y.E.; Chun, H.; Jo, S.; Bae, J.Y.; An, H.; Min, J.O.; Oh, S.J.; Han, K.S.; et al. Glial GABA, synthesized by monoamine oxidase B, mediates tonic inhibition. J. Physiol. 2014, 592, 4951–4968. [Google Scholar] [CrossRef] [PubMed]
  160. Mederos, S.; Sánchez-Puelles, C.; Esparza, J.; Valero, M.; Ponomarenko, A.; Perea, G. GABAergic signaling to astrocytes in the prefrontal cortex sustains goal-directed behaviors. Nat. Neurosci. 2021, 24, 82–92. [Google Scholar] [CrossRef] [PubMed]
  161. Jo, S.; Yarishkin, O.; Hwang, Y.J.; Chun, Y.E.; Park, M.; Woo, D.H.; Bae, J.Y.; Kim, T.; Lee, J.; Chun, H.; et al. GABA from reactive astrocytes impairs memory in mouse models of Alzheimer’s disease. Nat. Med. 2014, 20, 886–896. [Google Scholar] [CrossRef]
  162. Carter, C.L.; Resnick, E.M.; Mallampalli, M.; Kalbarczyk, A. Sex and Gender Differences in Alzheimer’s Disease: Recommendations for Future Research. J. Womens Health 2012, 21, 1018–1023. [Google Scholar] [CrossRef] [PubMed]
  163. Viña, J.; Lloret, A. Why Women Have More Alzheimer’s Disease Than Men: Gender and Mitochondrial Toxicity of Amyloid-β Peptide. J. Alzheimers Dis. 2010, 20, S527–S533. [Google Scholar] [CrossRef] [Green Version]
  164. Roy, U.; Stute, L.; Höfling, C.; Hartlage-Rübsamen, M.; Matysik, J.; Roβner, S.; Alia, A. Sex- and age-specific modulation of brain GABA levels in a mouse model of Alzheimer’s disease. Neurobiol. Aging 2018, 62, 168–179. [Google Scholar] [CrossRef] [PubMed]
  165. Hansen, K.B.; Yi, F.; Perszyk, R.E.; Furukawa, H.; Wollmuth, L.P.; Gibb, A.J.; Traynelis, S.F. Structure, function, and allosteric modulation of NMDA receptors. J. Gen. Physiol. 2018, 150, 1081–1105. [Google Scholar] [CrossRef]
  166. Panatier, A.; Theodosis, D.T.; Mothet, J.-P.; Touquet, B.; Pollegioni, L.; Poulain, D.A.; Oliet, S.H.R. Glia-Derived d-Serine Controls NMDA Receptor Activity and Synaptic Memory. Cell 2006, 125, 775–784. [Google Scholar] [CrossRef]
  167. Wolosker, H.; Balu, D.T. D-Serine as the gatekeeper of NMDA receptor activity: Implications for the pharmacologic management of anxiety disorders. Transl. Psychiatry 2020, 10, 184. [Google Scholar] [CrossRef] [PubMed]
  168. Van Horn, M.; Sild, M.; Ruthazer, E. D-serine as a gliotransmitter and its roles in brain development and disease. Front. Cell. Neurosci. 2013, 7, 39. [Google Scholar] [CrossRef] [Green Version]
  169. Madeira, C.; Lourenco, M.V.; Vargas-Lopes, C.; Suemoto, C.K.; Brandão, C.O.; Reis, T.; Leite, R.E.P.; Laks, J.; Jacob-Filho, W.; Pasqualucci, C.A.; et al. d-serine levels in Alzheimer’s disease: Implications for novel biomarker development. Transl. Psychiatry 2015, 5, e561. [Google Scholar] [CrossRef] [PubMed]
  170. Balu, D.T.; Pantazopoulos, H.; Huang, C.C.Y.; Muszynski, K.; Harvey, T.L.; Uno, Y.; Rorabaugh, J.M.; Galloway, C.R.; Botz-Zapp, C.; Berretta, S.; et al. Neurotoxic astrocytes express the d-serine synthesizing enzyme, serine racemase, in Alzheimer’s disease. Neurobiol. Dis. 2019, 130, 104511. [Google Scholar] [CrossRef] [PubMed]
  171. Piubelli, L.; Pollegioni, L.; Rabattoni, V.; Mauri, M.; Princiotta Cariddi, L.; Versino, M.; Sacchi, S. Serum d-serine levels are altered in early phases of Alzheimer’s disease: Towards a precocious biomarker. Transl. Psychiatry 2021, 11, 77. [Google Scholar] [CrossRef]
  172. Pajarillo, E.; Rizor, A.; Lee, J.; Aschner, M.; Lee, E. The role of astrocytic glutamate transporters GLT-1 and GLAST in neurological disorders: Potential targets for neurotherapeutics. Neuropharmacology 2019, 161, 107559. [Google Scholar] [CrossRef]
  173. Todd, A.C.; Hardingham, G.E. The Regulation of Astrocytic Glutamate Transporters in Health and Neurodegenerative Diseases. Int. J. Mol. Sci. 2020, 21, 9607. [Google Scholar] [CrossRef] [PubMed]
  174. Lewerenz, J.; Maher, P. Chronic Glutamate Toxicity in Neurodegenerative Diseases—What is the Evidence? Front. Neurosci. 2015, 9, 469. [Google Scholar] [CrossRef] [PubMed]
  175. Xing, L.; Yang, T.; Cui, S.; Chen, G. Connexin Hemichannels in Astrocytes: Role in CNS Disorders. Front. Mol. Neurosci. 2019, 12, 23. [Google Scholar] [CrossRef] [Green Version]
  176. Huang, X.; Su, Y.; Wang, N.; Li, H.; Li, Z.; Yin, G.; Chen, H.; Niu, J.; Yi, C. Astroglial Connexins in Neurodegenerative Diseases. Front. Mol. Neurosci. 2021, 14, 657514. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of a tripartite synapse. The tripartite synapse is composed of presynaptic and postsynaptic processes with astrocytic processes enwrapping the synapses. (1) The release of neurotransmitter from the presynaptic terminal acts on the postsynaptic terminal as well as with astrocytic receptors mediating intracellular calcium elevation via G-protein coupled receptors (GPCRs). (2) Then, calcium elevation triggers the release of gliotransmitters that bind with the postsynaptic terminal receptors (3) or presynaptic receptors (4) to modulate synaptic transmission.
Figure 1. Schematic representation of a tripartite synapse. The tripartite synapse is composed of presynaptic and postsynaptic processes with astrocytic processes enwrapping the synapses. (1) The release of neurotransmitter from the presynaptic terminal acts on the postsynaptic terminal as well as with astrocytic receptors mediating intracellular calcium elevation via G-protein coupled receptors (GPCRs). (2) Then, calcium elevation triggers the release of gliotransmitters that bind with the postsynaptic terminal receptors (3) or presynaptic receptors (4) to modulate synaptic transmission.
Ijms 22 07887 g001
Figure 2. Schematic representation of ATP and glutamate release from astrocytes in AD pathology. (1) Amyloid plaques react with P2Y1R or α7 nAChRs to evoke anomalous and increased intracellular calcium increases in astrocytes (2). These calcium increases induce the release of glutamate and/or ATP via vesicle release or hemichannel Cx43 release (3). ATP and glutamate act on P2X7Rs or NMDARs to increase the excitability and calcium concentrations of neurons (4). Arrows indicate increases.
Figure 2. Schematic representation of ATP and glutamate release from astrocytes in AD pathology. (1) Amyloid plaques react with P2Y1R or α7 nAChRs to evoke anomalous and increased intracellular calcium increases in astrocytes (2). These calcium increases induce the release of glutamate and/or ATP via vesicle release or hemichannel Cx43 release (3). ATP and glutamate act on P2X7Rs or NMDARs to increase the excitability and calcium concentrations of neurons (4). Arrows indicate increases.
Ijms 22 07887 g002
Figure 3. Schematic representation of GABA release from astrocytes in AD pathology. (1) Amyloid plaques evoke an enhanced activation of the monoamine oxidase MAOB (2) to induce the GABA synthesis using putrescine as a substrate (3). GABA is released from astrocytes via the anion channel BEST1 (4). GABA released as a gliotransmitter acts on postsynaptic (5) or presynaptic (6) GABA receptors (GABAR) to impact synaptic transmission. Arrows indicate increases.
Figure 3. Schematic representation of GABA release from astrocytes in AD pathology. (1) Amyloid plaques evoke an enhanced activation of the monoamine oxidase MAOB (2) to induce the GABA synthesis using putrescine as a substrate (3). GABA is released from astrocytes via the anion channel BEST1 (4). GABA released as a gliotransmitter acts on postsynaptic (5) or presynaptic (6) GABA receptors (GABAR) to impact synaptic transmission. Arrows indicate increases.
Ijms 22 07887 g003
Figure 4. Summary of astrocyte–neuron communication alterations in Alzheimer´s disease. Representation of an astrocytic and neuronal network in the normal state (left) and during the pathogenesis of AD (right). The magnification in the middle depicts a tripartite synapse in the normal state (left) and AD (right). Neurons are colored in green and astrocytes are colored in blue. Spontaneous astrocytic calcium signals are augmented in AD. The increase in calcium activity has been suggested to be mediated through upregulation of the purinergic receptor P2Y1, α7 nAChR, and/or the metabotropic glutamate receptor mGluR5 in astrocytes. Elevations in astrocytic intracellular calcium concentration promote the up-regulation of Cx43, leading to enhanced release of gliotransmitters (glutamate, ATP, and GABA). Diseased astrocytes express lower amounts of the glutamate transporter EAAT2 (or GLT-1), contributing to the accumulation of glutamate in the synaptic cleft. The increase in glutamate and ATP will overstimulate ionotropic receptors with consequent neuronal death. Thus, the presence of Aβ and the malfunctioning of astrocytes in AD lead to impairments in the neuronal network, decreasing the density of synaptic puncta and the basal synaptic transmission, which in turn will produce impairments in learning and memory. AD: Alzheimer´s disease; α7 nAChR: α7 nicotinic acetylcholine receptors; Cx43: connexin 43; eNMDAR: extra-synaptic NMDA receptor.
Figure 4. Summary of astrocyte–neuron communication alterations in Alzheimer´s disease. Representation of an astrocytic and neuronal network in the normal state (left) and during the pathogenesis of AD (right). The magnification in the middle depicts a tripartite synapse in the normal state (left) and AD (right). Neurons are colored in green and astrocytes are colored in blue. Spontaneous astrocytic calcium signals are augmented in AD. The increase in calcium activity has been suggested to be mediated through upregulation of the purinergic receptor P2Y1, α7 nAChR, and/or the metabotropic glutamate receptor mGluR5 in astrocytes. Elevations in astrocytic intracellular calcium concentration promote the up-regulation of Cx43, leading to enhanced release of gliotransmitters (glutamate, ATP, and GABA). Diseased astrocytes express lower amounts of the glutamate transporter EAAT2 (or GLT-1), contributing to the accumulation of glutamate in the synaptic cleft. The increase in glutamate and ATP will overstimulate ionotropic receptors with consequent neuronal death. Thus, the presence of Aβ and the malfunctioning of astrocytes in AD lead to impairments in the neuronal network, decreasing the density of synaptic puncta and the basal synaptic transmission, which in turn will produce impairments in learning and memory. AD: Alzheimer´s disease; α7 nAChR: α7 nicotinic acetylcholine receptors; Cx43: connexin 43; eNMDAR: extra-synaptic NMDA receptor.
Ijms 22 07887 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nanclares, C.; Baraibar, A.M.; Araque, A.; Kofuji, P. Dysregulation of Astrocyte–Neuronal Communication in Alzheimer’s Disease. Int. J. Mol. Sci. 2021, 22, 7887. https://doi.org/10.3390/ijms22157887

AMA Style

Nanclares C, Baraibar AM, Araque A, Kofuji P. Dysregulation of Astrocyte–Neuronal Communication in Alzheimer’s Disease. International Journal of Molecular Sciences. 2021; 22(15):7887. https://doi.org/10.3390/ijms22157887

Chicago/Turabian Style

Nanclares, Carmen, Andres Mateo Baraibar, Alfonso Araque, and Paulo Kofuji. 2021. "Dysregulation of Astrocyte–Neuronal Communication in Alzheimer’s Disease" International Journal of Molecular Sciences 22, no. 15: 7887. https://doi.org/10.3390/ijms22157887

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