Next Article in Journal
Influence of Tamoxifen on Different Biological Pathways in Tumorigenesis and Transformation in Adipose-Derived Stem Cells, Mammary Cells and Mammary Carcinoma Cell Lines—An In Vitro Study
Next Article in Special Issue
Increased KV2.1 Channel Clustering Underlies the Reduction of Delayed Rectifier K+ Currents in Hippocampal Neurons of the Tg2576 Alzheimer’s Disease Mouse
Previous Article in Journal
Manufacture and Quality Control of Human Umbilical Cord-Derived Mesenchymal Stem Cell Sheets for Clinical Use
Previous Article in Special Issue
CaMKII Inhibition Attenuates Distinct Gain-of-Function Effects Produced by Mutant Nav1.6 Channels and Reduces Neuronal Excitability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Microglia and Astrocytes in Alzheimer’s Disease: From Neuroinflammation to Ca2+ Homeostasis Dysregulation

by
Giulia Di Benedetto
,
Chiara Burgaletto
,
Carlo Maria Bellanca
,
Antonio Munafò
,
Renato Bernardini
* and
Giuseppina Cantarella
Department of Biomedical and Biotechnological Sciences, Section of Pharmacology, University of Catania, 95123 Catania, Italy
*
Author to whom correspondence should be addressed.
Cells 2022, 11(17), 2728; https://doi.org/10.3390/cells11172728
Submission received: 8 August 2022 / Revised: 26 August 2022 / Accepted: 30 August 2022 / Published: 1 September 2022

Abstract

:
Alzheimer’s disease (AD) is the most common form of dementia worldwide, with a complex, poorly understood pathogenesis. Cerebral atrophy, amyloid-β (Aβ) plaques, and neurofibrillary tangles represent the main pathological hallmarks of the AD brain. Recently, neuroinflammation has been recognized as a prominent feature of the AD brain and substantial evidence suggests that the inflammatory response modulates disease progression. Additionally, dysregulation of calcium (Ca2+) homeostasis represents another early factor involved in the AD pathogenesis, as intracellular Ca2+ concentration is essential to ensure proper cellular and neuronal functions. Although growing evidence supports the involvement of Ca2+ in the mechanisms of neurodegeneration-related inflammatory processes, scant data are available on its contribution in microglia and astrocytes functioning, both in health and throughout the AD continuum. Nevertheless, AD-related aberrant Ca2+ signalling in astrocytes and microglia is crucially involved in the mechanisms underpinning neuroinflammatory processes that, in turn, impact neuronal Ca2+ homeostasis and brain function. In this light, we attempted to provide an overview of the current understanding of the interactions between the glia cells-mediated inflammatory responses and the molecular mechanisms involved in Ca2+ homeostasis dysregulation in AD.

Graphical Abstract

1. Introduction

Alzheimer’s disease (AD) is an aggressive neurodegenerative disorder and the leading cause of dementia in the elderly population worldwide, characterized by impairment of cognitive functions including memory, decision-making, orientation to physical surroundings and language [1,2]. The acknowledged neuropathological hallmarks of AD include extracellular senile plaques, composed of amyloid-beta (Aβ) peptides, and intracellular neurofibrillary tangles (NFTs) generated by the hyper-phosphorylated form of the tau protein, as well as selective neuronal and synaptic loss in specific brain regions.
To date, despite the extensive research efforts, there are no disease-modifying therapeutic options and the currently available Food and Drug Administration (FDA) approved pharmacotherapy for AD provide only a temporary symptomatic relief, without affecting pathophysiological disease progression. Due to these results, researchers are attempting multifaceted and integrated approaches, ranging from drug repurposing to stem cell technologies [3,4]. While compelling evidence suggests that AD has a multifactorial aetiology [5], a robust glial-mediated inflammatory response manifests as an early feature of AD pathophysiological mechanisms and plays a pivotal etiopathogenic role [6,7], due to the ability to exacerbate Aβ and Tau pathologies [8,9]. The inflammatory response in AD acts as a “double-edged sword”. In fact, while on the one hand, during the initial phases of AD, neuroinflammation represents a self-defence mechanism that protects the brain, by promoting tissue repair and quick clearance of noxious stimuli; on the other, as the disease progresses, sustained inflammatory response brings about detrimental effects, fuelling neurodegeneration [10,11]. Despite the neuroinflammatory events being not exclusively limited to the brain, as peripheral immune dysfunctions occur early in the disease, perhaps even prior to relevant pathogenic cerebral changes [12,13], glial cells, especially microglia and astrocytes, make up the brain’s innate immune system, engage in a fine-tuned and tight crosstalk, and represent the major players in the ongoing neuroinflammatory response in AD [14]. Remarkably, a high number of reactive astrocytes and activated microglia are found near the senile plaques of individuals with AD, suggesting their crucial involvement in AD pathogenesis [15,16,17]. Nevertheless, it is also known that an aberrant calcium (Ca2+) signalling represents another hallmark of AD and is reported not only in neuronal cells, but also in astrocytes and microglia [18,19]. Glial cells profoundly contribute to the inflammatory response that, in turn, impacts on neuronal Ca2+ homeostasis and brain function [14], leading to age-related cognitive decline. Indeed, intracellular Ca2+ is a second messenger that regulates many facets of neuronal physiology, including neurotransmitter release, growth, differentiation, and synaptic plasticity, as well as pathophysiological processes, such apoptosis, necrosis, and degeneration [20].
Ca2+ dysregulation and neuroinflammation exhibit an extensive and tight crosstalk, leading to distinct outcomes in different neural cell types [21]. Ca2+ dysregulation in glial cells, with special regard to astrocytes, enhances the activity of Ca2+ dependent protein phosphatase calcineurin (CaN), which shows elevated activity in animal models of aging and disease, resulting in increased production and release of several inflammatory mediators, such as cytokines (i.e., interleukin (IL)-1β and tumour necrosis factor-α (TNF-α) [21]. In turn, pro-inflammatory cytokines, produced and released by astrocytes and microglia, sustain activation of the MAP kinase signalling pathways (p38), and the expression/activity of L-type voltage sensitive Ca2+ channels in neurons, causing Ca2+ dysregulation, hyperactive calcineurin activity, and synaptic plasticity deficits [21,22] (Figure 1).
In such scenario, we focused on the crucial involvement of glia cell-mediated inflammatory responses and calcium homeostasis dysregulation in AD, with the aim to identify new and promising molecular targets for the development of innovative therapies for this disease.

2. Neuroinflammatory Scenario in Alzheimer’s Disease: The Role of Microglia and Astrocytes

2.1. Microglia and Astrocytes: The Main Actors of the Neuroinflammatory Response

Over the years, several hypotheses have been proposed to unveil the complex pathological mechanisms underlying AD-related neurodegeneration [23]. However, the ultimate aetiology of AD remains obscure.
The discovery of elevated levels of inflammatory markers associated with cognitive decline in AD patients and the identification of AD risk genes associated with innate immune functions suggest that neuroinflammation holds a crucial role in AD pathogenesis [24]. Neuroinflammation not only occurs in an early phase of the disease, but may also drive its progression [25].
The main immune cell components of the central nervous system (CNS) microglia and astrocytes, represent the cellular master regulators of inflammatory responses in the brain [10], and as resident innate immune cells, are strongly implicated in aberrant molecular pathways that underlie AD pathogenetic alterations [26]. Traditionally, both types of cells can assume multiple reactive phenotypes, neurotoxic and neuroprotective, which are related to the stage of the disease and to the regional location [10].
Such complexity in their distribution could be one of the reasons why clinical trials of anti-inflammatory therapies, to date, have failed to improve clinical outcomes of AD patients [27]. In this light, a better understanding of the roles of microglia and astrocytes in AD is essential for developing effective therapies.

2.2. Microglia: The Sentinels of the Brain

Microglia cells, the brain tissue macrophages, represent slightly less than 10% of CNS cells. Nevertheless, these innate immune cells appear as the principal responders to both immunological and neuronal stimuli. In fact, microglial cells are involved in homeostasis and immune surveillance of the brain: they act as “sentinels” detecting changes in the surrounding environment by means of a cluster of genes termed “sensome” [28], migrate to injured sites and protect against noxious stimuli, producing cytokines and chemokines to recruit additional cells and remove pathological agents [10,29,30]. Microglia are also active players in complex neurodevelopmental programs such as neurogenesis and synaptic pruning [31]. In the adult brain, microglia interact with neurons and macroglia cells to provide trophic support, synaptic modulation, and neuronal reorganization [32,33]. Loss of homeostasis or tissue changes occurring in AD, induce various dynamic microglial processes, including changes in cell morphology (ramified or quiescent/resting, activated or ameboid), surface phenotype and proliferative responses [34].
Microglia have a complex role in the trajectory of AD owing to their diverse phe-notypes and variety of activation pathways [24]. In fact, the role of microglia in neurodegeneration related to AD evolves with the progression of the disease, revealing the complexity of the factors regulating the balance between its beneficial and detrimental effects. Microglia express a family of innate immune cell receptors, the pattern recognition receptors (PRRs) (i.e., receptors for advanced glycation end (RAGE), Toll-like receptors (TLRs), and scavenger receptors). Different Aβ species, including neurotoxic oligomeric Aβ, can activate microglia by binding to PRRs, suggesting that accumulating Aβ, is responsible for the neuroinflammation observed in AD [35,36]. In the early disease stages, the activation of microglia has been associated with neuroprotective effects by removing the harmful stimuli represented by hyper-production of Aβ, clearing cell debris, phagocytosing dead cells, and releasing neurotrophic factors [37]. However, the persistence of the damaging stimuli through the disease continuum, subjects these cells to a state of chronic activation and inflammatory cytokines release, that drives neurotoxicity and neurodegeneration. Danger-associated molecular patterns (DAMPs), that arise from these processes, further activate microglia initiating a self-perpetuating proinflammatory cascade that leads to Aβ clearance ability [29,38]. In addition, microglia seem to be intimately involved also in the propagation of tau pathology; however, it is still unclear whether microglia contribute to tau pathology by failing to engulf it or by releasing factors that exacerbate tau pathology [34]. Thus, identifying the role of disease-specific microglial phenotypes during AD progression may be of interest in designing immunotherapy that boost or temper inflammation depending on the disease stage.

2.3. Astrocyte: A Crucial Regulator of Brain Homeostasis

Astrocytes are the most represented glial cells in the CNS [39]. They participate in neurovascular coupling and in the integrity maintenance of the blood–brain barrier (BBB), regulate cerebral blood flow, and mediate synaptic formation and transmission [40]. Astrocytes constitute the parenchymal part of the BBB and regulate the establishment of endothelial cell-to-cell junctions that preserve the structural and functional integrity of the BBB through secretion of soluble factors, such as growth factors (i.e., glial cell-line neurotrophic factor (GDNF), vascular endothelial growth factor (VEGF), basic fibroblast growth factor (bFGF), angiopoetin-1 (ANG-1), morphogens (Sonic hedgehog (Shh) and Wnt) and extracellular vesicles [41,42].
The astrocyte–endotheliocyte communication is crucial in various neurological diseases associated with BBB impairments. In physiological circumstances, perivascular astrocytic end-feet are not tightly sealed [41]. On the contrary, in neuroinflammatory conditions related to AD, and characterized by BBB breakdown and leukocyte infiltration into the CNS, end-feet of reactive astrocytes mount a parenchymal line of defence by inducing tight junctions formation in response to inflammatory cues, thereby tightening the border to limit brain infiltration of peripheral immunocyte [43]. In this scenario, astrocytes activity may exacerbate inflammatory reactions and tissue damage, or promote immunosuppression and tissue repair depending on timing and disease context [44].
Astrocytes also exert neuroprotective function in AD, blunting the plaque build-up by Aβ clearance [45]. Conversely, astrocytes are induced by noxious stimuli, Aβ or activated microglia. Activated microglia release specific astrocyte-activating signals, such as interlukin-1 alpha (IL-1α), complement component 1q (C1q) and TNF-α, which, in turn, activate β-secretase and γ-secretase activity, cleaving APP, and stimulating β-amyloid formation by astrocytes, thereby supplementing neuronal β-amyloid production [26,46]. Therefore, inflammatory stimulation of astrocytes can induce increased expression of β-site APP-cleaving enzyme (BACE1), APP, and β-secretase, implying a feed-forward mechanism of astrocytic Aβ production [47,48].

2.4. Microglia and Astrocytes Activation Profile in Alzheimer’s Disease Is Heterogeneous

Although traditionally classified as either proinflammatory (neurotoxic), or immunoregulatory (neuroprotective), the phenotypic distribution of microglia and astrocytes can change based on the progression of the disease [10]. Microglia and astrocytes become activated around the senile plaques in AD brain, and the resulting morphological changes are typical of an Aβ-induced neuroinflammatory response that, in turn, induce neurodegeneration [17,49,50]. Consistently with this evidence, glial activation and subsequent inflammatory events are not only a simple response to Aβ deposition, but rather represent a key contributor to AD pathogenesis [51]. Thus, a deep understanding of the precise role of glial cells during disease progression appears essential for developing novel and effective therapies.

2.5. The Neuroinflammatory Cycle: Microglia and Astrocytes Crosstalk in Alzheimer’s Disease and Potential Therapeutic Approaches

Neuroinflammation is the result of a complex of responses generated by a network of interactions between cells, such as neurons, microglia, astrocytes, and brain infiltrating leukocytes [26]. Particularly, microglia and astrocyte reactions encompass pro- and anti-inflammatory responses which attempt to defend neurons from damaging stimuli (i.e., pathogens and misfolded proteins), simultaneously promoting neuronal repair. Moreover, glial activity represents a self-limiting defence reaction aimed to contain and eradicate threats from the host organism [52]. Despite this, aging and neurodegenerative disorders such as AD determine functional alterations of microglia and astrocytes, subjecting these cells to a state of chronic activation. Persistent inflammation evolves when the initial acute inflammatory response is not adequately resolutive, leading to the accumulation of TNF, IFNs, and IL-6, as well as immunocytes, apoptotic cells, and debris at the site of insult/injury, thus compromising functional homeostasis [7].
Moreover, astrocytes and microglia release several signalling molecules (i.e., growth factors, neurotransmitters, cytokines, chemokines) establishing a bidirectional crosstalk for a tight reciprocal modulation to support neuronal survival and function during CNS insult or injury [33,53].
Specifically, upon injury or disease, damaged neurons release self-antigens or modified proteins that activate resting microglia. The latter, as the major immunological effector of the innate immune system in the brain and first line of defence, migrate to sites of damage to engulf dead cells and debris [54]. When inflammatory/pathogenetic stimuli persist, microglia remain trapped in a vicious cycle, characterized by a state of “non-functional” chronic activation. Activated microglia also release astrocyte-activating signals, including IL-1α, TNF-α, TNF-related apoptosis-inducing ligand (TRAIL) and C1q, which induce neuroinflammatory astrocytes that, in turn, amplify the neurodegenerative cycle [26,55].
The above reported bidirectional interactions set the stage for an endless cycle accelerating disease progression which could represent a relevant target to open new avenues into devising the long-due, disease-modifying therapies for AD.
Current evidence generally supports a simplified model of neuroinflammation in which the chronic inflammatory background provides an initial mild stimulus that results in glial priming, tied to a possible protective wave for such activated glia in the early stages of AD, along with the emergence of Aβ burden [24]. Nonetheless, the ineffective clearance of Aβ, combined with tau aggregation, impair these defence properties, and elicit an ongoing harmful glial activation processes over the course of AD [56].
Nevertheless, the huge plethora of different functions exerted by glial cells, combined with the available evidence suggesting the presence of distinct patterns of activation and the lack of a proper knowledge regarding the molecular regulations underlying cellular transitions, delay the development and design of pharmacological approaches targeting these specific glial subtypes [57,58]. This aim could be pursued through different ways, including the specific suppression of glial pro-inflammatory properties and the modulation of phenotypic changes in order to counteract microglial priming and to maintain protective and phagocytic properties for prolonged periods [59].
Several preclinical studies and phase I/II clinical trials assessing the targeting of Triggering Receptor Expressed on Myeloid cells 2 (TREM2) [60], NOD-like receptor protein 3 (NLRP3) [61], TLRs [62,63], or CD33 [64] have shown encouraging results. However, though promising, these potential therapeutic approaches are still in their early phases and further efforts are needed to develop a deeper understanding regarding the tight relationship between the mechanistic and regulatory functions of different astrocytic and microglial subtypes and their precise contributions in different stages of AD progression.

3. The Relevance of Ca2+ Homeostasis on Microglia and Astrocytes

3.1. Intracellular Ca2+ Concentration ([Ca2+]i): A Key Read-Out of Glial Activity

Evidence demonstrates that glial cells undergo complex changes in intracellular Ca2+ concentration ([Ca2+]i), which are considered a key read-out of glial activity and reactivity [65,66]. Glial [Ca2+]i is four to five orders of magnitude less than that in the narrow system of clefts that constitutes the extracellular environment of the CNS [67,68]. Indeed, a steep electrochemical gradient supports the entry of Ca2+, causing a transient increase in cell membrane permeability to Ca2+ [69]. Glial cells exhibit a dynamical regulation of intra- and intercellular content of Ca2+, both in their steady state and in response to specific stimuli [70]. In fact, Ca2+ stores within the cell may release Ca2+ in response to specific intracellular messengers, increasing [Ca2+]i [71]. Such transient rises of [Ca2+]i, due to the presence of Ca2+-permeable membrane ion channels, regulate several intracellular events, such as metabolic processes, gene expression, and ion transport systems [72]. Moreover, changes in [Ca2+]i can be considered as a second messenger system which coordinate the changes occurring in the extracellular environment with intracellular processes [73].

3.2. The Versatility of Ca2+ as an Essential Driver of Biological Processes in CNS

Glial cells express a variety of ion channels, and neurotransmitter transporters and receptors, that enable them to detect and respond to neuronal and synaptic activity [74,75]. In physiological conditions, receptor-mediated stimuli or electrical pulses in neurons generate different Ca2+ signals, with distinct spatial dimensions, amplitude, oscillations, temporal extension, and subcellular localization [76,77]. Spatial variability ranges from nanodomains in cellular organelles, up to gradients which cover the whole cell body. Changes in glial [Ca2+]i have been measured under various conditions, as glial cells are responding to chemical, mechanical and electrical stimuli [78]. Fluctuations in [Ca2+]i are not only a passive response, but seem also correlated with the reaction of glial cells to environmental changes, which, in turn, lead to a modification in glial function; therefore, these are not passive responses and can be rather considered as a form of glial excitability Ca2+-mediated [76,77]. The versatility of Ca2+ as an intracellular messenger is derived from different cytosolic Ca2+ concentrations, generating by regulated openings of Ca2+-permeable channels expressed both in plasma membrane and in different organelles [71]. Indeed, it is known that the activation of microglial cells increases Ca2+ permeability, which leads to Ca2+ influx into the cell [79,80]. Such an increase in [Ca2+]i acts as a second messenger that initiates signalling cascades, driving to essential biologic processes, such as cell proliferation, differentiation, migration, as well as activation of intracellular enzymatic pathways involved in transcriptional regulation of various genes [81].
Several intracellular organelles contain Ca2+ concentrations higher than the surrounding cytoplasm, acting as Ca2+ stores, but, in most mammalian cells, the major intracellular Ca2+ pool is the endoplasmic reticulum (ER) [71]. In microglial cells, the release of Ca2+ from the intracellular stores is controlled both by Ca2+ itself and by messengers, such as inositol-1,4,5-triphosphate (IP3) [81]. The stimulation of different receptors leads to the activation of phospholipase C (PLC), which hydrolyze phosphoinositide-4,5-bisphosphate (PIP2) to IP3 and 1,2-diacyl-glycerol (DAG). IP3 in turn induces the release of Ca2+ from intracellular stores, and DAG is associated with the entry of Ca2+ from the extracellular compartment. The IP3 interacts with its receptor (IP3R) on the ER walls, leading to the release of stored Ca2+ into the cytosol [82]. Another receptor for Ca2+ release located in the membrane of the ER is the ryanodine receptor (RyR) [83]. RyRs are regulated by multiple factors, including Ca2+ itself and the intracellular messenger cyclic adenosine diphosphate ribose (cADPR) [84]. In vitro studies on cultured human microglia from both foetal and adult brain specimens revealed the expression of mRNA for RyR1 and RyR2, whereas RyR3 mRNA was detected only in foetal microglia [85].
Besides IP3, the two other intracellular second messengers, cADPR and nicotinic acid adenine dinucleotide phosphate (NAADP), produced by the ADP-ribosyl cyclase CD38, are also involved in the mobilization of Ca2+ from intracellular stores [86]. RyRs are cADPR targets, whereas the receptor for NAADP is still unknown. NAADP could serve as a regulator for RyRs, two-pore channels, and transient receptor potential channel, subtype mucolipin (TRP-ML) channels [87,88]. The functional roles of cADPR have been well studied in microglia. In fact, several studies report that cADPR promotes microglial activation and secretion of proinflammatory cytokines [86,89,90]. On the other hand, little is known of NAADP and further studies are needed to clarify the role of NAADP in the regulation of Ca2+ homeostasis in microglial cells [88].

3.3. Mechanisms Mediating Ca2+ Exit from the Microglial Cytosol: The Role of Pumps and Exchangers

There is ample evidence showing that Ca2+ entering microglial cells through various routes, has to be finely tuned, as the accumulation of intracellular Ca2+ would lead to cell dysfunctions [79,91]. Mechanisms mediating Ca2+ exit from the cytosol involve various pumps and exchangers on the plasma membrane [92]. They include Na+/Ca2+ exchangers (NCX), Na+/Ca2+-K+ pumps (NCKX), and plasma-membrane Ca2+ ATPase pumps (PMCAs) [93,94]. In contrast to the abundancy of reports of NCX functions in microglial cells, little information is available on the properties of microglial Ca2+ pumps. It is known that NCX can operate either in the forward mode, coupling the uphill extrusion of Ca2+ (three ions) to the influx of Na+ (one ion), or in the reverse mode, coupling the extrusion of Na+ to the influx of Ca2+ ions [95,96]. Currently, three NCX genes (NCX1, NCX2, and NCX3), have been identified; in addition, it was reported that NCX1 is the most represented subtype in microglia [97].

3.4. Astrocyte Excitability Depends upon Variations of Cytosolic Ca2+ Concentration

Among glial cells, astrocytes were first characterized as non-excitable cells of the central nervous system, although expressing voltage-gated channels [98]. Astrocyte excitability results, in fact, from variations of cytosolic Ca2+ concentration [99]. At the cellular level, those Ca2+ signals occur in astrocytes in response to synaptic activity and may cause release of molecules such as glutamate, ATP, TNF-α, or D-serine, which modulate synaptic transmission [100,101,102,103], as well as vasoconstriction/vasodilatation [104,105,106,107]. The tight association of astrocytes to pre- and post- synaptic elements, both structurally and functionally, is referred to as tripartite synapse [108,109,110]. In addition, astrocytic Ca2+ signals can modulate neuronal synchronization and firing pattern [111,112,113] and have been observed in vivo, in response to external stimuli [114,115]. Astrocytes possess a complex Ca2+ signalling machinery that relies on several Ca2+ mobilizing pathways associated with ER Ca2+ release and plasmalemmal Ca2+ entry through ion channels and the reversed NCX [116].
The remarkable heterogeneity of astroglial Ca2+ signalling is strictly associated to the extensive adaptive potential of astrocytes, which may tailor Ca2+ signalling toolkits to meet a multitude of challenges. The identification of targets for the Ca2+ signals in the physiological and the pathological context remains a compelling task.

4. The Role of Ca2+ Homeostasis Dysfunction in Microglia and Astrocytes in Alzheimer’s Disease

Glial–calcium signalling acts as a molecular mechanism for integration within glial syncytium and between glial and neuronal circuits [117]. Quite naturally, this powerful signalling system also plays an important role in neuropathology, such as AD [68,69] (Figure 2). The highly dynamic intra- and intercellular Ca2+ signalling of astrocytes and microglia is quickly perturbed under neuropathological conditions, determining the change in cellular response and transformation into reactive phenotypes [118,119,120].

4.1. Microglial Ca2+ Signaling in Alzheimer’s Disease

Microglial cells from AD patients exhibit significant abnormalities in Ca2+-mediated signals compared with non-demented brains [19,121]. In fact, it has been reported that stimulation of microglia cultured with Aβ25–35 peptide results in a transient increase in intracellular Ca2+ [122,123]. Many microglia functions are mediated by Ca2+ ions [124,125,126]. In particular, microglia reaction is accompanied with intracellular Ca2+ increase, a process that is required to induce the release of cytokines and chemokines [127]. McLarnon et al. [19], have showed significant perturbations in Ca2+-mediated signal transduction (i.e., global elevation of basal levels of Ca2+, rapid depletion of Ca2+ from endoplasmic reticulum, diminished response to ATP stimulation) in microglia from AD patients.
Moreover, an increased Ca2+ response to a purinergic P2X7 receptor (P2X7R) agonist was found in cultured foetal human microglial cells exposed to Aβ1-42 and, consistently, enhanced P2X7R expression was stated in AD microglia [128,129]. Aβ-dependent mitochondrial toxicity in mouse microglia requires P2X7R expression and is prevented by nimodipine, a potent L-type calcium channel (LTCC) inhibitor [130].
Aberrant activation of the NLRP3 inflammasome was observed in AD; therefore, fine tuning of its activity appears essential for maintenance of proper cellular homeostasis [131,132,133,134]. NLRP3 knockout, or a dysfunction of its key signalling components significantly reduces Aβ-induced microglial activation in vitro [135], decreases Aβ deposition or Tau pathology, and also alleviates cognitive impairment in AD mouse models [136,137]. Aβ increases [Ca2+]i levels, which, in turn, contribute to the activation of the NLRP3 inflammasome in microglia [130]. Lee et al. [138] reported that the Ca2+-sensing receptor (CaSR) activates the NLRP3 inflammasome by increasing [Ca2+]i levels and decreasing cyclic adenosine 3′,5′-monophosphate (cAMP) levels.
Recently, the role of Ca2+ homeostasis modulator family proteins (Calhm, Calhm1, Calhm2, and Calhm3) has gained increasing attention in the field of AD research [139]. Calhm1 controls Ca2+ homeostasis, Aβ production, and neuronal cell vulnerability to Aβ-induced toxicity. Moreover, the P86L mutation of Calhm1 is correlated with the incidence of AD [140]. Calhm2, which is highly expressed in the murine brain, could regulate ATP release in astrocytes and its deletion induces a depression-like phenotype in mice [141]. In a recent study, it was shown that the expression of Calhm2 is increased in microglia of an AD mouse model [142]. In particular, in 5×FAD mice carrying five familial AD gene mutations, both conventional knockout of Calhm2 and conditional microglial knockout of Calhm2 significantly reduce Aβ deposition, neuroinflammation, and cognitive impairments. Knockout of Calhm2 inhibited microglial proinflammatory activity, but increases phagocytic activity, leading to restoration of the balance between inflammation and phagocytosis. Moreover, knockout of Calhm2 reduces acute LPS-induced neuroinflammation [142].
In addition, changes in properties of store-operated Ca2+ entry in microglia could lead to altered immune cell response and neurovascular unit dysfunction in the inflamed AD brain [143]. The membrane protein TREM2 regulates key microglial functions including phagocytosis and chemotaxis. Loss of function variants of TREM2 are associated with increased risk of AD [144]. Loss of TREM2 confers a defect in microglial Ca2+ response to purinergic signals, suggesting a window of Ca2+ signalling for optimal microglial motility [145]. An increase in microglial [Ca2+]i is linked to phosphorylation of cAMP response element-binding protein (CREB) and the activation of the pCREB pathway. In turn, CREB phosphorylation leads, for example, to the microglial secretion of proinflammatory cytokines, such as IL-6, and other molecules such as (Brain-derived neurotrophic factor) BDNF [125].

4.2. Astrocytic Dysregulation of Ca2+ Homeostasis in Alzheimer’s Disease

Recent studies have reported the existence of a causal relationship between astrocytes and neuroinflammation [10]. Alteration of Ca2+ homeostasis in astrocytes has been regarded as an important component of AD [146]. Aβ disrupt Ca2+ homeostasis, specifically in hippocampal astrocytes by targeting key elements of Ca2+ signalling, such as mGluR5 and IP3R1 [147]. Consistently, several studies reported that Aβ triggered transient [Ca2+]i increases or oscillations in astrocytes [148,149,150].
Takano et al. [151] reported that after the injection of Aβ in both Dutch/Iowa mice and control mice, the frequency of astrocytic Ca2+ oscillations increased significantly and became approximately three times more than the control group.
Generally, Ca2+ entry into astrocytes includes active transport by different types of voltage gated calcium channels (VGCCs) distributed in the membrane and passive leakage [152]. Upregulation of L-type VGCCs expression in astrocytes has been shown in pathological conditions, such as those related to AD [153]. Anekonda et al. [154] have demonstrated that L-type Ca2+ current blockers can protect cells from the inductive effect of Aβ. Ahmadpour et al. [152] confirmed the efficacy of blocking VGCCs, but the recovery was not significant enough.
Aβ has been shown to impair the membrane permeability, due to the Aβ-formed pores resulting in an increase in Ca2+ influx through the membrane [155]. This additional influx leads to damage to astrocytic homeostasis in simulations. Especially, when oscillations disappear and the astrocyte Ca2+ level is at a high steady state, Aβ can cause dramatic increases in [Ca2+]i. In fact, channel formation has been already proposed as a molecular mechanism for Aβ toxicity in the early 1990s [156].
Although disruptions on the membrane by Aβ are believed to be an important mechanism, the intracellular signalling pathways also deserve attention. In particular, the toxic effects of Aβ increase the channel opening probability of RyRs [157]. In fact, several studies also suggest that Aβ can directly increase the RyR expression [158,159]. In addition, Oulès et al. [160] have reported that RyR-mediated Ca2+ release can be reduced after treatment with RyR inhibitor.

4.3. How Does Aβ Disrupt Ca2+ Homeostasis?

Since it is well known that Aβ can disrupt RyR-regulated Ca2+ signals, the mechanisms for how this happens remains controversial [161,162,163]. In fact, the influence of Aβ on individual pumps, channels, and exchangers remains largely unknown [164,165]. Although it is unclear how Aβ disrupts intracellular Ca2+ homeostasis, there is growing evidence that Aβ directly affects the production of IP3 [166,167], calcium-induced calcium release through RyR [168,169], and the induction of Ca2+-permeability across membrane pores [170,171].
Notably, it was reported that the accumulation of Aβ oligomers can trigger aberrant intracellular Ca2+ levels by disrupting the intrinsic Ca2+ regulatory mechanism within cells [172,173]. These disruptions can cause changes in homeostasis levels that can have detrimental effects on cell function and survival. Although studies have shown that Aβ can interfere with Ca2+ fluxes, the complexity of these interactions remains elusive [174,175]. The formation of Aβ plaques and fibrils are a consequence of the imbalance between the formation and sequestration of Aβ [176,177]. Consequently, the slow accumulation of Aβ peptides can alter Ca2+ signalling processes within neurons and glial cells, leading to synaptic failure and neuronal death [178,179,180]. Moreover, the accumulation of Aβ has been linked to the progression of AD by altering Ca2+ signalling processes, but the underlying mechanisms for how and why this occurs are not fully understood [167,181].
The above-reported data indicate that Ca2+ dysregulation in AD occurs as a result of hyperactivity of Ca2+ channels in the plasma membrane and in intracellular compartments. Moreover, it seems to be restricted not only to neurons, but rather it represents a global phenomenon that affects many cell types in the brain. These observations suggest that the identification of molecules which affect Ca2+ homeostasis dysfunction in microglia and astrocytes could be a promising new therapeutic target in the treatment of AD.

5. Conclusions

Persistent and self-fuelling inflammatory processes represent the major driving force that characterize the progressive nature of AD. Microglia and astrocytes represent critical actors of neuroinflammatory response, as their interaction play a crucial role at different stages of brain pathology.
Sentinel microglia are highly active surveyors of their microenvironment, and upon detecting pathogens or damage, become activated via the common molecule signals (i.e., TNF-α, IL-1α, and C1α) to trigger reactive astrocytes. Astrocytes amplify neuroinflammation, due to their function and strategic position to mobilize peripheral immunity.
The aberrant glial cell biology seen in AD might be involved in the disruption of Ca2+ homeostasis in neurons, which negatively impacts on functioning and viability of neuronal cells, leading to negative functional outcomes and cognitive decline. As glia cell-mediated inflammatory responses and Ca2+ dysregulation in AD share a common language, it appears reasonable to hypothesize a crucial role of the latter in the neurodegenerative processes associated with dementia.
In such scenario, targeting Ca2+-glia-AD axis may result beneficial for AD therapy in terms of delayed disease progression but, given the limited amount of evidence currently available, further investigations are needed to determine the paths of all actors in this promising field of research.

Author Contributions

Literature survey, writing—original draft preparation, G.D.B., C.B., A.M. and C.M.B.; writing—review and editing, G.D.B., C.B., G.C. and R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the PRIN grant no. 2017YH3SXK from the Italian Ministry of Research.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Giulia Di Benedetto is a Researcher at the University of Catania within the EU-funded PON REACT project (Azione IV.6–“Contratti di ricerca su tematiche green” del nuovo Asse IV del PON Ricerca e Innovazione 2014–2020 “Istruzione e ricerca per il recupero–REACT–EU”; Progetto “Green chemistry: dagli scarti agroalimentari a molecole bioattive nel trattamento delle malattie neurodegenerative”, CUP E65F21002560001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Scheltens, P.; De Strooper, B.; Kivipelto, M.; Holstege, H.; Chételat, G.; Teunissen, C.E.; Cummings, J.; van der Flier, W.M. Alzheimer’s disease. Lancet 2021, 397, 1577–1590. [Google Scholar] [CrossRef]
  2. Monfared, A.A.T.; Byrnes, M.J.; White, L.A.; Zhang, Q. Alzheimer’s Disease: Epidemiology and Clinical Progression. Neurol. Ther. 2022, 11, 553–569. [Google Scholar] [CrossRef] [PubMed]
  3. Munafò, A.; Burgaletto, C.; Di Benedetto, G.; Di Mauro, M.; Di Mauro, R.; Bernardini, R.; Cantarella, G. Repositioning of Immunomodulators: A Ray of Hope for Alzheimer’s Disease? Front. Neurosci. 2020, 14, 614643. [Google Scholar] [CrossRef]
  4. Bonaventura, G.; Munafò, A.; Bellanca, C.; La Cognata, V.; Iemmolo, R.; Attaguile, G.; Di Mauro, R.; Di Benedetto, G.; Cantarella, G.; Barcellona, M.; et al. Stem Cells: Innovative Therapeutic Options for Neurodegenerative Diseases? Cells 2021, 10, 1992. [Google Scholar] [CrossRef]
  5. Gong, C.-X.; Liu, F.; Iqbal, K. Multifactorial Hypothesis and Multi-Targets for Alzheimer’s Disease. J. Alzheimer’s Dis. 2018, 64, S107–S117. [Google Scholar] [CrossRef] [PubMed]
  6. Heppner, F.L.; Ransohoff, R.M.; Becher, B. Immune attack: The role of inflammation in Alzheimer disease. Nat. Rev. Neurosci. 2015, 16, 358–372. [Google Scholar] [CrossRef]
  7. Onyango, I.; Jauregui, G.; Čarná, M.; Bennett, J.; Stokin, G. Neuroinflammation in Alzheimer’s Disease. Biomedicines 2021, 9, 524. [Google Scholar] [CrossRef]
  8. Carret-Rebillat, A.-S.; Pace, C.; Gourmaud, S.; Ravasi, L.; Montagne-Stora, S.; Longueville, S.; Tible, M.; Sudol, E.; Chang, R.C.-C.; Paquet, C.; et al. Neuroinflammation and Aβ Accumulation Linked To Systemic Inflammation Are Decreased By Genetic PKR Down-Regulation. Sci. Rep. 2015, 5, srep08489. [Google Scholar] [CrossRef]
  9. Metcalfe, M.; Figueiredo-Pereira, M.E. Relationship Between Tau Pathology and Neuroinflammation in Alzheimer’s Disease. Mt. Sinai J. Med. A J. Transl. Pers. Med. 2010, 77, 50–58. [Google Scholar] [CrossRef] [PubMed]
  10. Kwon, H.S.; Koh, S.-H. Neuroinflammation in neurodegenerative disorders: The roles of microglia and astrocytes. Transl. Neurodegener. 2020, 9, 42. [Google Scholar] [CrossRef]
  11. Kaur, D.; Sharma, V.; Deshmukh, R. Activation of microglia and astrocytes: A roadway to neuroinflammation and Alzheimer’s disease. Inflammopharmacology 2019, 27, 663–677. [Google Scholar] [CrossRef]
  12. Dá Mesquita, S.; Ferreira, A.C.; Sousa, J.C.; Correia-Neves, M.; Sousa, N.; Marques, F. Insights on the pathophysiology of Alzheimer’s disease: The crosstalk between amyloid pathology, neuroinflammation and the peripheral immune system. Neurosci. Biobehav. Rev. 2016, 68, 547–562. [Google Scholar] [CrossRef]
  13. Burgaletto, C.; Platania, C.B.M.; Di Benedetto, G.; Munafò, A.; Giurdanella, G.; Federico, C.; Caltabiano, R.; Saccone, S.; Conti, F.; Bernardini, R.; et al. Targeting the miRNA-155/TNFSF10 network restrains inflammatory response in the retina in a mouse model of Alzheimer’s disease. Cell Death Dis. 2021, 12, 905. [Google Scholar] [CrossRef]
  14. Brawek, B.; Garaschuk, O. Network-wide dysregulation of calcium homeostasis in Alzheimer’s disease. Cell Tissue Res. 2014, 357, 427–438. [Google Scholar] [CrossRef]
  15. Zotova, E.; Holmes, C.; Johnston, D.; Neal, J.W.; Nicoll, J.A.R.; Boche, D. Microglial alterations in human Alzheimer’s disease following Aβ42 immunization. Neuropathol. Appl. Neurobiol. 2011, 37, 513–524. [Google Scholar] [CrossRef]
  16. Jung, S.; Schwartz, M. Non-Identical Twins—Microglia and Monocyte-Derived Macrophages in Acute Injury and Autoimmune Inflammation. Front. Immunol. 2012, 3, 89. [Google Scholar] [CrossRef]
  17. Moreira, E.G.; Boll, K.M.; Correia, D.G.; Soares, J.F.; Rigobello, C.; Maes, M. Why Should Psychiatrists and Neuroscientists Worry about Paraoxonase 1? Curr. Neuropharmacol. 2019, 17, 1004–1020. [Google Scholar] [CrossRef]
  18. Acosta, C.; Anderson, H.D.; Anderson, C.M. Astrocyte dysfunction in Alzheimer disease. J. Neurosci. Res. 2017, 95, 2430–2447. [Google Scholar] [CrossRef]
  19. McLarnon, J.G.; Choi, H.B.; Lue, L.-F.; Walker, D.G.; Kim, S.U. Perturbations in calcium-mediated signal transduction in microglia from Alzheimer’s disease patients. J. Neurosci. Res. 2005, 81, 426–435. [Google Scholar] [CrossRef]
  20. Yu, J.-T.; Chang, R.C.-C.; Tan, L. Calcium dysregulation in Alzheimer’s disease: From mechanisms to therapeutic opportunities. Prog. Neurobiol. 2009, 89, 240–255. [Google Scholar] [CrossRef]
  21. Sama, D.M.; Norris, C.M. Calcium dysregulation and neuroinflammation: Discrete and integrated mechanisms for age-related synaptic dysfunction. Ageing Res. Rev. 2013, 12, 982–995. [Google Scholar] [CrossRef]
  22. Price, B.R.; Johnson, L.A.; Norris, C.M. Reactive astrocytes: The nexus of pathological and clinical hallmarks of Alzheimer’s disease. Ageing Res. Rev. 2021, 68, 101335. [Google Scholar] [CrossRef]
  23. Tam, K.Y.; Ju, Y. Pathological mechanisms and therapeutic strategies for Alzheimer’s disease. Neural Regen. Res. 2022, 17, 543. [Google Scholar] [CrossRef]
  24. Leng, F.; Edison, P. Neuroinflammation and microglial activation in Alzheimer disease: Where do we go from here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef]
  25. 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]
  26. Burgaletto, C.; Munafò, A.; Di Benedetto, G.; De Francisci, C.; Caraci, F.; Di Mauro, R.; Bucolo, C.; Bernardini, R.; Cantarella, G. The immune system on the TRAIL of Alzheimer’s disease. J. Neuroinflammation 2020, 17, 298. [Google Scholar] [CrossRef]
  27. Imbimbo, B.P.; Solfrizzi, V.; Panza, F. Are NSAIDs useful to treat Alzheimer’s disease or mild cognitive impairment? Front. Aging Neurosci. 2010, 2, 19. [Google Scholar] [CrossRef]
  28. E Hickman, S.; Kingery, N.D.; Ohsumi, T.K.; Borowsky, M.L.; Wang, L.-C.; Means, T.K.; El Khoury, J. The microglial sensome revealed by direct RNA sequencing. Nat. Neurosci. 2013, 16, 1896–1905. [Google Scholar] [CrossRef]
  29. Hickman, S.; Izzy, S.; Sen, P.; Morsett, L.; El Khoury, J. Microglia in neurodegeneration. Nat. Neurosci. 2018, 21, 1359–1369. [Google Scholar] [CrossRef]
  30. Glass, C.K.; Saijo, K.; Winner, B.; Marchetto, M.C.; Gage, F.H. Mechanisms Underlying Inflammation in Neurodegeneration. Cell 2010, 140, 918–934. [Google Scholar] [CrossRef] [Green Version]
  31. Cowan, M.; Petri, W.A.J. Microglia: Immune Regulators of Neurodevelopment. Front. Immunol. 2018, 9, 2576. [Google Scholar] [CrossRef] [PubMed]
  32. Schafer, D.P.; Lehrman, E.K.; Kautzman, A.G.; Koyama, R.; Mardinly, A.R.; Yamasaki, R.; Ransohoff, R.M.; Greenberg, M.E.; Barres, B.A.; Stevens, B. Microglia sculpt postnatal neural circuits in an activity and complement-dependent manner. Neuron 2012, 74, 691–705. [Google Scholar] [CrossRef]
  33. Matejuk, A.; Ransohoff, R.M. Crosstalk Between Astrocytes and Microglia: An Overview. Front. Immunol. 2020, 11, 1416. [Google Scholar] [CrossRef] [PubMed]
  34. Sarlus, H.; Heneka, M.T. Microglia in Alzheimer’s disease. J. Clin. Investig. 2017, 127, 3240–3249. [Google Scholar] [CrossRef] [PubMed]
  35. Venegas, C.; Heneka, M.T. Danger-associated molecular patterns in Alzheimer’s disease. J. Leukoc. Biol. 2016, 101, 87–98. [Google Scholar] [CrossRef]
  36. Yang, T.; Li, S.; Xu, H.; Walsh, D.M.; Selkoe, D.J. Large Soluble Oligomers of Amyloid β-Protein from Alzheimer Brain Are Far Less Neuroactive Than the Smaller Oligomers to Which They Dissociate. J. Neurosci. 2016, 37, 152–163. [Google Scholar] [CrossRef]
  37. Simon, E.; Obst, J.; Gomez-Nicola, D. The Evolving Dialogue of Microglia and Neurons in Alzheimer’s Disease: Microglia as Necessary Transducers of Pathology. Neuroscience 2019, 405, 24–34. [Google Scholar] [CrossRef]
  38. Ransohoff, R.M.; El Khoury, J. Microglia in Health and Disease. Cold Spring Harb. Perspect. Biol. 2015, 8, a020560. [Google Scholar] [CrossRef]
  39. Placone, A.L.; McGuiggan, P.M.; Bergles, D.; Guerrero-Cazares, H.; Quiñones-Hinojosa, A.; Searson, P.C. Human astrocytes develop physiological morphology and remain quiescent in a novel 3D matrix. Biomaterials 2014, 42, 134–143. [Google Scholar] [CrossRef]
  40. Patabendige, A.; Singh, A.; Jenkins, S.; Sen, J.; Chen, R. Astrocyte Activation in Neurovascular Damage and Repair Following Ischaemic Stroke. Int. J. Mol. Sci. 2021, 22, 4280. [Google Scholar] [CrossRef]
  41. Pivoriūnas, A.; Verkhratsky, A. Astrocyte–Endotheliocyte Axis in the Regulation of the Blood–Brain Barrier. Neurochem. Res. 2021, 46, 2538–2550. [Google Scholar] [CrossRef] [PubMed]
  42. Alvarez, J.I.; Katayama, T.; Prat, A. Glial influence on the blood brain barrier. Glia 2013, 61, 1939–1958. [Google Scholar] [CrossRef] [PubMed]
  43. Quintana, F.J. Astrocytes to the rescue! Glia limitans astrocytic endfeet control CNS inflammation. J. Clin. Investig. 2017, 127, 2897–2899. [Google Scholar] [CrossRef]
  44. Colombo, E.; Farina, C. Astrocytes: Key Regulators of Neuroinflammation. Trends Immunol. 2016, 37, 608–620. [Google Scholar] [CrossRef]
  45. Joly-Amado, A.; Hunter, J.; Quadri, Z.; Zamudio, F.; Rocha-Rangel, P.V.; Chan, D.; Kesarwani, A.; Nash, K.; Lee, D.C.; Morgan, D.; et al. CCL2 Overexpression in the Brain Promotes Glial Activation and Accelerates Tau Pathology in a Mouse Model of Tauopathy. Front. Immunol. 2020, 11, 997. [Google Scholar] [CrossRef] [PubMed]
  46. Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.-S.; Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017, 541, 481–487. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, J.; O’Connor, T.; Vassar, R. The contribution of activated astrocytes to Aβ production: Implications for Alzheimer’s disease pathogenesis. J. Neuroinflammation 2011, 8, 150. [Google Scholar] [CrossRef]
  48. Monterey, M.D.; Wei, H.; Wu, X.; Wu, J.Q. The Many Faces of Astrocytes in Alzheimer’s Disease. Front. Neurol. 2021, 12, 619626. [Google Scholar] [CrossRef]
  49. Bouvier, D.S.; Jones, E.V.; Quesseveur, G.; Davoli, M.A.; Ferreira, T.; Quirion, R.; Mechawar, N.; Murai, K.K. High Resolution Dissection of Reactive Glial Nets in Alzheimer’s Disease. Sci. Rep. 2016, 6, 24544. [Google Scholar] [CrossRef]
  50. Perez-Nievas, B.G.; Serrano-Pozo, A. Deciphering the Astrocyte Reaction in Alzheimer’s Disease. Front. Aging Neurosci. 2018, 10, 114. [Google Scholar] [CrossRef] [Green Version]
  51. Vandenbark, A.A.; Offner, H.; Matejuk, S.; Matejuk, A. Microglia and astrocyte involvement in neurodegeneration and brain cancer. J. Neuroinflammation 2021, 18, 298. [Google Scholar] [CrossRef] [PubMed]
  52. Carson, M.J.; Thrash, C.; Walter, B. The cellular response in neuroinflammation: The role of leukocytes, microglia and astrocytes in neuronal death and survival. Clin. Neurosci. Res. 2006, 6, 237–245. [Google Scholar] [CrossRef]
  53. Jha, M.K.; Jo, M.; Kim, J.-H.; Suk, K. Microglia-Astrocyte Crosstalk: An Intimate Molecular Conversation. Neurosci. 2018, 25, 227–240. [Google Scholar] [CrossRef] [PubMed]
  54. Hanisch, U.-K.; Kettenmann, H. Microglia: Active sensor and versatile effector cells in the normal and pathologic brain. Nat. Neurosci. 2007, 10, 1387–1394. [Google Scholar] [CrossRef] [PubMed]
  55. Cantarella, G.; Lempereur, L.; D’Alcamo, M.A.; Risuglia, N.; Cardile, V.; Pennisi, G.; Scoto, G.M.; Bernardini, R. Trail interacts redundantly with nitric oxide in rat astrocytes: Potential contribution to neurodegenerative processes. J. Neuroimmunol. 2007, 182, 41–47. [Google Scholar] [CrossRef]
  56. Guerriero, F.; Sgarlata, C.; Francis, M.; Maurizi, N.; Faragli, A.; Perna, S.; Rondanelli, M.; Rollone, M.; Ricevuti, G. Neuroinflammation, immune system and Alzheimer disease: Searching for the missing link. Aging Clin. Exp. Res. 2017, 29, 821–831. [Google Scholar] [CrossRef]
  57. Li, T.; Lu, L.; Pember, E.; Li, X.; Zhang, B.; Zhu, Z. New Insights into Neuroinflammation Involved in Pathogenic Mechanism of Alzheimer’s Disease and Its Potential for Therapeutic Intervention. Cells 2022, 11, 1925. [Google Scholar] [CrossRef]
  58. Dhapola, R.; Hota, S.S.; Sarma, P.; Bhattacharyya, A.; Medhi, B.; Reddy, D.H. Recent advances in molecular pathways and therapeutic implications targeting neuroinflammation for Alzheimer’s disease. Inflammopharmacology 2021, 29, 1669–1681. [Google Scholar] [CrossRef]
  59. Hampel, H.; Caraci, F.; Cuello, A.C.; Caruso, G.; Nisticò, R.; Corbo, M.; Baldacci, F.; Toschi, N.; Garaci, F.; Chiesa, P.A.; et al. A Path Toward Precision Medicine for Neuroinflammatory Mechanisms in Alzheimer’s Disease. Front. Immunol. 2020, 11, 456. [Google Scholar] [CrossRef]
  60. Price, B.R.; Sudduth, T.L.; Weekman, E.M.; Johnson, S.; Hawthorne, D.; Woolums, A.; Wilcock, D.M. Therapeutic Trem2 activation ameliorates amyloid-beta deposition and improves cognition in the 5XFAD model of amyloid deposition. J. Neuroinflammation 2020, 17, 238. [Google Scholar] [CrossRef]
  61. Coll, R.C.; Robertson, A.A.B.; Chae, J.J.; Higgins, S.C.; Muñoz-Planillo, R.; Inserra, M.C.; Vetter, I.; Dungan, L.S.; Monks, B.G.; Stutz, A.; et al. A small-molecule inhibitor of the NLRP3 inflammasome for the treatment of inflammatory diseases. Nat. Med. 2015, 21, 248–255. [Google Scholar] [CrossRef] [PubMed]
  62. Cui, W.; Sun, C.; Ma, Y.; Wang, S.; Wang, X.; Zhang, Y. Inhibition of TLR4 Induces M2 Microglial Polarization and Provides Neuroprotection via the NLRP3 Inflammasome in Alzheimer’s Disease. Front. Neurosci. 2020, 14, 444. [Google Scholar] [CrossRef] [PubMed]
  63. Rubio-Araiz, A.; Finucane, O.M.; Keogh, S.; Lynch, M. Anti-TLR2 antibody triggers oxidative phosphorylation in microglia and increases phagocytosis of β-amyloid. J. Neuroinflammation 2018, 15, 247. [Google Scholar] [CrossRef]
  64. Griciuc, A.; Federico, A.N.; Natasan, J.; Forte, A.M.; McGinty, D.; Nguyen, H.; Volak, A.; LeRoy, S.; Gandhi, S.; Lerner, E.P.; et al. Gene therapy for Alzheimer’s disease targeting CD33 reduces amyloid beta accumulation and neuroinflammation. Hum. Mol. Genet. 2020, 29, 2920–2935. [Google Scholar] [CrossRef] [PubMed]
  65. Verkhratsky, A. Calcium signalling in glial cells. Neurophysiology 1997, 29, 205–212. [Google Scholar] [CrossRef]
  66. Finkbeiner, S.M. Glial calcium. Glia 1993, 9, 83–104. [Google Scholar] [CrossRef]
  67. Zündorf, G.; Reiser, G. Calcium Dysregulation and Homeostasis of Neural Calcium in the Molecular Mechanisms of Neurodegenerative Diseases Provide Multiple Targets for Neuroprotection. Antioxidants Redox Signal. 2011, 14, 1275–1288. [Google Scholar] [CrossRef]
  68. Lim, D.; Ronco, V.; Grolla, A.A.; Verkhratsky, A.; Genazzani, A.A. Glial Calcium Signalling in Alzheimer’s Disease. Rev. Physiol. Biochem. Pharmacol. 2014, 167, 45–65. [Google Scholar] [CrossRef]
  69. Ureshino, R.P.; Erustes, A.G.; Bassani, T.B.; Wachilewski, P.; Guarache, G.C.; Nascimento, A.C.; Costa, A.J.; Smaili, S.S.; Pereira, G.J.D.S. The Interplay between Ca2+ Signaling Pathways and Neurodegeneration. Int. J. Mol. Sci. 2019, 20, 6004. [Google Scholar] [CrossRef]
  70. Rieder, P.; Gobbo, D.; Stopper, G.; Welle, A.; Damo, E.; Kirchhoff, F.; Scheller, A. Astrocytes and Microglia Exhibit Cell-Specific Ca2+ Signaling Dynamics in the Murine Spinal Cord. Front. Mol. Neurosci. 2022, 15, 840948. [Google Scholar] [CrossRef]
  71. Bagur, R.; Hajnóczky, G. Intracellular Ca2+ Sensing: Its Role in Calcium Homeostasis and Signaling. Mol. Cell 2017, 66, 780–788. [Google Scholar] [CrossRef] [PubMed]
  72. Stutzmann, G.E.; Mattson, M.P. Endoplasmic Reticulum Ca2+ Handling in Excitable Cells in Health and Disease. Pharmacol. Rev. 2011, 63, 700–727. [Google Scholar] [CrossRef] [PubMed]
  73. Verkhratsky, A.; Parpura, V. Calcium signalling and calcium channels: Evolution and general principles. Eur. J. Pharmacol. 2013, 739, 1–3. [Google Scholar] [CrossRef]
  74. Perea, G.; Sur, M.; Araque, A. Neuron-glia networks: Integral gear of brain function. Front. Cell. Neurosci. 2014, 8, 378. [Google Scholar] [CrossRef]
  75. Araque, A.; Navarrete, M. Glial cells in neuronal network function. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 2375–2381. [Google Scholar] [CrossRef]
  76. Verkhratsky, A.; Kettenmann, H. Calcium signalling in glial cells. Trends Neurosci. 1996, 19, 346–352. [Google Scholar] [CrossRef]
  77. Vernadakis, A. GLIA-NEURON INTERCOMMUNICATIONS AND SYNAPTIC PLASTICITY. Prog. Neurobiol. 1996, 49, 185–214. [Google Scholar] [CrossRef]
  78. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef]
  79. Mizoguchi, Y.; Monji, A. Microglial Intracellular Ca2+ Signaling in Synaptic Development and its Alterations in Neurodevelopmental Disorders. Front. Cell. Neurosci. 2017, 11, 69. [Google Scholar] [CrossRef] [PubMed]
  80. Sharma, P.; Ping, L. Calcium ion influx in microglial cells: Physiological and therapeutic significance. J. Neurosci. Res. 2014, 92, 409–423. [Google Scholar] [CrossRef]
  81. Berridge, M.J. The Inositol Trisphosphate/Calcium Signaling Pathway in Health and Disease. Physiol. Rev. 2016, 96, 1261–1296. [Google Scholar] [CrossRef]
  82. Krebs, J.; Agellon, L.B.; Michalak, M. Ca2+ homeostasis and endoplasmic reticulum (ER) stress: An integrated view of calcium signaling. Biochem. Biophys. Res. Commun. 2015, 460, 114–121. [Google Scholar] [CrossRef]
  83. Ruiz, A.; Matute, C.; Alberdi, E. Intracellular Ca2+ release through ryanodine receptors contributes to AMPA receptor-mediated mitochondrial dysfunction and ER stress in oligodendrocytes. Cell Death Dis. 2010, 1, e54. [Google Scholar] [CrossRef] [PubMed]
  84. Ogunbayo, O.A.; Zhu, Y.; Rossi, D.; Sorrentino, V.; Ma, J.; Zhu, M.X.; Evans, A.M. Cyclic Adenosine Diphosphate Ribose Activates Ryanodine Receptors, whereas NAADP Activates Two-pore Domain Channels. J. Biol. Chem. 2011, 286, 9136–9140. [Google Scholar] [CrossRef] [PubMed]
  85. Klegeris, A.; Choi, H.B.; McLarnon, J.G.; McGeer, P.L. Functional ryanodine receptors are expressed by human microglia and THP-1 cells: Their possible involvement in modulation of neurotoxicity. J. Neurosci. Res. 2007, 85, 2207–2215. [Google Scholar] [CrossRef]
  86. Franco, L.; Bodrato, N.; Moreschi, I.; Usai, C.; Bruzzone, S.; Scarfì, S.; Zocchi, E.; De Flora, A. Cyclic ADP-ribose is a second messenger in the lipopolysaccharide-stimulated activation of murine N9 microglial cell line. J. Neurochem. 2006, 99, 165–176. [Google Scholar] [CrossRef] [PubMed]
  87. Guse, A.H. Linking NAADP to Ion Channel Activity: A Unifying Hypothesis. Sci. Signal. 2012, 5, pe18. [Google Scholar] [CrossRef]
  88. Galione, A. NAADP Receptors. Cold Spring Harb. Perspect. Biol. 2019, 11, a035071. [Google Scholar] [CrossRef]
  89. Mayo, L.; Jacob-Hirsch, J.; Amariglio, N.; Rechavi, G.; Moutin, M.-J.; Lund, F.E.; Stein, R. Dual Role of CD38 in Microglial Activation and Activation-Induced Cell Death. J. Immunol. 2008, 181, 92–103. [Google Scholar] [CrossRef]
  90. Bodrato, N.; Franco, L.; Fresia, C.; Guida, L.; Usai, C.; Salis, A.; Moreschi, I.; Ferraris, C.; Verderio, C.; Basile, G.; et al. Abscisic Acid Activates the Murine Microglial Cell Line N9 through the Second Messenger Cyclic ADP-ribose. J. Biol. Chem. 2009, 284, 14777–14787. [Google Scholar] [CrossRef] [Green Version]
  91. Brawek, B.; Schwendele, B.; Riester, K.; Kohsaka, S.; Lerdkrai, C.; Liang, Y.; Garaschuk, O. Impairment of in vivo calcium signaling in amyloid plaque-associated microglia. Acta Neuropathol. 2014, 127, 495–505. [Google Scholar] [CrossRef]
  92. Brini, M.; Carafoli, E. The Plasma Membrane Ca2+ ATPase and the Plasma Membrane Sodium Calcium Exchanger Cooperate in the Regulation of Cell Calcium. Cold Spring Harb. Perspect. Biol. 2010, 3, a004168. [Google Scholar] [CrossRef]
  93. Stafford, N.; Wilson, C.; Oceandy, D.; Neyses, L.; Cartwright, E.J. The Plasma Membrane Calcium ATPasesand Their Role as Major New Players in Human Disease. Physiol. Rev. 2017, 97, 1089–1125. [Google Scholar] [CrossRef]
  94. Wu, X.; Weng, L.; Zhang, J.; Liu, X.; Huang, J. The Plasma Membrane Calcium ATPases in Calcium Signaling Network. Curr. Protein Pept. Sci. 2018, 19, 813–822. [Google Scholar] [CrossRef] [PubMed]
  95. Molinaro, P.; Pannaccione, A.; Sisalli, M.J.; Secondo, A.; Cuomo, O.; Sirabella, R.; Cantile, M.; Ciccone, R.; Scorziello, A.; di Renzo, G.; et al. A New Cell-penetrating Peptide That Blocks the Autoinhibitory XIP Domain of NCX1 and Enhances Antiporter Activity. Mol. Ther. 2015, 23, 465–476. [Google Scholar] [CrossRef]
  96. Di Benedetto, G.; Valerio, O.; Lariccia, V.; Burgaletto, C.; Lempereur, L.; Parenti, C.; Zanghì, G.N.; Matteucci, A.; Amoroso, S.; Bernardini, R.; et al. Tumor necrosis factor-related apoptosis-inducing ligand reduces the expression of the neuroprotective Na+/Ca2+ exchanger isoform NCX 3 in human neuroblastoma SH—SY 5Y cells. FEBS J. 2018, 286, 737–749. [Google Scholar] [CrossRef]
  97. Newell, E.W.; Stanley, E.F.; Schlichter, L.C. Reversed Na+/Ca2+ Exchange Contributes to Ca2+ Influx and Respiratory Burst in Microglia. Channels 2007, 1, 366–376. [Google Scholar] [CrossRef]
  98. Sontheimer, H. Voltage-dependent ion channels in glial cells. Glia 1994, 11, 156–172. [Google Scholar] [CrossRef] [PubMed]
  99. Cornell-Bell, A.H.; Finkbeiner, S.M.; Cooper, M.S.; Smith, S.J. Glutamate Induces Calcium Waves in Cultured Astrocytes: Long-Range Glial Signaling. Science 1990, 247, 470–473. [Google Scholar] [CrossRef]
  100. 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]
  101. Gómez-Gonzalo, M.; Navarrete, M.; Perea, G.; Covelo, A.; Martín-Fernández, M.; Shigemoto, R.; Luján, R.; Araque, A. Endocannabinoids Induce Lateral Long-Term Potentiation of Transmitter Release by Stimulation of Gliotransmission. Cereb. Cortex 2014, 25, 3699–3712. [Google Scholar] [CrossRef] [PubMed]
  102. Panatier, A.; Theodosis, D.T.; Mothet, J.-P.; Touquet, B.; Pollegioni, L.; Poulain, D.A.; Oliet, S.H. Glia-Derived d-Serine Controls NMDA Receptor Activity and Synaptic Memory. Cell 2006, 125, 775–784. [Google Scholar] [CrossRef] [PubMed]
  103. Takata, N.; Mishima, T.; Hisatsune, C.; Nagai, T.; Ebisui, E.; Mikoshiba, K.; Hirase, H. Astrocyte Calcium Signaling Transforms Cholinergic Modulation to Cortical Plasticity In Vivo. J. Neurosci. 2011, 31, 18155–18165. [Google Scholar] [CrossRef] [PubMed]
  104. Metea, M.R. Glial Cells Dilate and Constrict Blood Vessels: A Mechanism of Neurovascular Coupling. J. Neurosci. 2006, 26, 2862–2870. [Google Scholar] [CrossRef]
  105. Mulligan, S.J.; MacVicar, B. Calcium transients in astrocyte endfeet cause cerebrovascular constrictions. Nat. 2004, 431, 195–199. [Google Scholar] [CrossRef] [PubMed]
  106. Takano, T.; Tian, G.-F.; Peng, W.; Lou, N.; Libionka, W.; Han, X.; Nedergaard, M. Astrocyte-mediated control of cerebral blood flow. Nat. Neurosci. 2005, 9, 260–267. [Google Scholar] [CrossRef]
  107. Zonta, M.; Sebelin, A.; Gobbo, S.; Fellin, T.; Pozzan, T.; Carmignoto, G. Glutamate-mediated cytosolic calcium oscillations regulate a pulsatile prostaglandin release from cultured rat astrocytes. J. Physiol. 2003, 553, 407–414. [Google Scholar] [CrossRef]
  108. Perea, G.; Navarrete, M.; Araque, A. Tripartite synapses: Astrocytes process and control synaptic information. Trends Neurosci. 2009, 32, 421–431. [Google Scholar] [CrossRef]
  109. Fiacco, T.A.; McCarthy, K.D. Multiple Lines of Evidence Indicate That Gliotransmission Does Not Occur under Physiological Conditions. J. Neurosci. 2018, 38, 3–13. [Google Scholar] [CrossRef]
  110. Savtchouk, I.; Volterra, A. Gliotransmission: Beyond Black-and-White. J. Neurosci. 2018, 38, 14–25. [Google Scholar] [CrossRef] [Green Version]
  111. 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] [PubMed]
  112. Perea, G. Properties of Synaptically Evoked Astrocyte Calcium Signal Reveal Synaptic Information Processing by Astrocytes. J. Neurosci. 2005, 25, 2192–2203. [Google Scholar] [CrossRef] [PubMed]
  113. Morquette, P.; Verdier, D.; Kadala, A.; Féthière, J.; Philippe, A.G.; Robitaille, R.; Kolta, A. An astrocyte-dependent mechanism for neuronal rhythmogenesis. Nat. Neurosci. 2015, 18, 844–854. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, X.; Lou, N.; Xu, Q.; Tian, G.-F.; Peng, W.G.; Han, X.; Kang, J.; Takano, T.; Nedergaard, M. Astrocytic Ca2+ signaling evoked by sensory stimulation in vivo. Nat. Neurosci. 2006, 9, 816–823. [Google Scholar] [CrossRef]
  115. Asada, A.; Ujita, S.; Nakayama, R.; Oba, S.; Ishii, S.; Matsuki, N.; Ikegaya, Y. Subtle modulation of ongoing calcium dynamics in astrocytic microdomains by sensory inputs. Physiol. Rep. 2015, 3, e12454. [Google Scholar] [CrossRef] [PubMed]
  116. Verkhratsky, A.; Semyanov, A.; Zorec, R. Physiology of Astroglial Excitability. Function 2020, 1, zqaa016. [Google Scholar] [CrossRef]
  117. Verkhratsky, A.; Anderova, M.; Chvatal, A. Differential calcium signalling in neuronal-glial networks. Front. Biosci. 2009, 14, 2004–2016. [Google Scholar] [CrossRef]
  118. Shigetomi, E.; Saito, K.; Sano, F.; Koizumi, S. Aberrant Calcium Signals in Reactive Astrocytes: A Key Process in Neurological Disorders. Int. J. Mol. Sci. 2019, 20, 996. [Google Scholar] [CrossRef]
  119. Korvers, L.; Costa, A.D.A.; Mersch, M.; Matyash, V.; Kettenmann, H.; Semtner, M. Spontaneous Ca2+ transients in mouse microglia. Cell Calcium 2016, 60, 396–406. [Google Scholar] [CrossRef]
  120. Heo, D.K.; Lim, H.M.; Nam, J.H.; Lee, M.G.; Kim, J.Y. Regulation of phagocytosis and cytokine secretion by store-operated calcium entry in primary isolated murine microglia. Cell. Signal. 2015, 27, 177–186. [Google Scholar] [CrossRef]
  121. McLarnon, J.G. Purinergic mediated changes in Ca2+ mobilization and functional responses in microglia: Effects of low levels of ATP. J. Neurosci. Res. 2005, 81, 349–356. [Google Scholar] [CrossRef] [PubMed]
  122. Combs, C.K.; Johnson, D.E.; Cannady, S.B.; Lehman, T.M.; Landreth, G.E. Identification of Microglial Signal Transduction Pathways Mediating a Neurotoxic Response to Amyloidogenic Fragments of β-Amyloid and Prion Proteins. J. Neurosci. 1999, 19, 928–939. [Google Scholar] [CrossRef] [PubMed]
  123. Hemonnot, A.-L.; Hua, J.; Ulmann, L.; Hirbec, H. Microglia in Alzheimer Disease: Well-Known Targets and New Opportunities. Front. Aging Neurosci. 2019, 11, 233. [Google Scholar] [CrossRef]
  124. Färber, K.; Kettenmann, H. Functional role of calcium signals for microglial function. Glia 2006, 54, 656–665. [Google Scholar] [CrossRef]
  125. Pan, K.; Garaschuk, O. The role of intracellular calcium-store-mediated calcium signals in in vivo sensor and effector functions of microglia. J. Physiol. 2022, in press. [Google Scholar] [CrossRef] [PubMed]
  126. Gehrmann, J.; Matsumoto, Y.; Kreutzberg, G.W. Microglia: Intrinsic immuneffector cell of the brain. Brain Res. Rev. 1995, 20, 269–287. [Google Scholar] [CrossRef]
  127. Hoffmann, A.; Kann, O.; Ohlemeyer, C.; Hanisch, U.-K.; Kettenmann, H. Elevation of Basal Intracellular Calcium as a Central Element in the Activation of Brain Macrophages (Microglia): Suppression of Receptor-Evoked Calcium Signaling and Control of Release Function. J. Neurosci. 2003, 23, 4410–4419. [Google Scholar] [CrossRef]
  128. McLarnon, J.G.; Ryu, M.J.K.; Walker, D.G.; Choi, B.H.B. Upregulated Expression of Purinergic P2X7 Receptor in Alzheimer Disease and Amyloid-β Peptide-Treated Microglia and in Peptide-Injected Rat Hippocampus. J. Neuropathol. Exp. Neurol. 2006, 65, 1090–1097. [Google Scholar] [CrossRef]
  129. Sanz, J.M.; Chiozzi, P.; Ferrari, D.; Colaianna, M.; Idzko, M.; Falzoni, S.; Fellin, R.; Trabace, L.; Di Virgilio, F. Activation of Microglia by Amyloid β Requires P2X7 Receptor Expression. J. Immunol. 2009, 182, 4378–4385. [Google Scholar] [CrossRef]
  130. Chiozzi, P.; Sarti, A.C.; Sanz, J.M.; Giuliani, A.L.; Adinolfi, E.; Vultaggio-Poma, V.; Falzoni, S.; Di Virgilio, F. Amyloid β-dependent mitochondrial toxicity in mouse microglia requires P2X7 receptor expression and is prevented by nimodipine. Sci. Rep. 2019, 9, 6475. [Google Scholar] [CrossRef] [Green Version]
  131. Zhang, Y.; Dong, Z.; Song, W. NLRP3 inflammasome as a novel therapeutic target for Alzheimer’s disease. Signal Transduct. Target. Ther. 2020, 5, 37. [Google Scholar] [CrossRef] [PubMed]
  132. Voet, S.; Srinivasan, S.; Lamkanfi, M.; Van Loo, G. Inflammasomes in neuroinflammatory and neurodegenerative diseases. EMBO Mol. Med. 2019, 11, e10248. [Google Scholar] [CrossRef] [PubMed]
  133. Van Zeller, M.; Dias, D.M.; Sebastião, A.M.; Valente, C.A. NLRP3 Inflammasome: A Starring Role in Amyloid-β- and Tau-Driven Pathological Events in Alzheimer’s Disease. J. Alzheimer’s Dis. 2021, 83, 939–961. [Google Scholar] [CrossRef] [PubMed]
  134. Sharma, B.R.; Kanneganti, T.-D. NLRP3 inflammasome in cancer and metabolic diseases. Nat. Immunol. 2021, 22, 550–559. [Google Scholar] [CrossRef]
  135. Halle, A.; Hornung, V.; Petzold, G.C.; Stewart, C.R.; Monks, B.G.; Reinheckel, T.; Fitzgerald, K.A.; Latz, E.; Moore, K.J.; Golenbock, D.T. The NALP3 inflammasome is involved in the innate immune response to amyloid-β. Nat. Immunol. 2008, 9, 857–865. [Google Scholar] [CrossRef]
  136. Ising, C.; Venegas, C.; Zhang, S.; Scheiblich, H.; Schmidt, S.V.; Vieira-Saecker, A.; Schwartz, S.; Albasset, S.; McManus, R.M.; Tejera, D.; et al. NLRP3 inflammasome activation drives tau pathology. Nature 2019, 575, 669–673. [Google Scholar] [CrossRef]
  137. Heneka, M.T.; Kummer, M.P.; Stutz, A.; Delekate, A.; Schwartz, S.; Vieira-Saecker, A.; Griep, A.; Axt, D.; Remus, A.; Tzeng, T.-C.; et al. NLRP3 is activated in Alzheimer’s disease and contributes to pathology in APP/PS1 mice. Nature 2013, 493, 674–678. [Google Scholar] [CrossRef]
  138. Lee, G.-S.; Subramanian, N.; Kim, A.I.; Aksentijevich, I.; Goldbach-Mansky, R.; Sacks, D.; Germain, R.N.; Kastner, D.L.; Chae, J.J. The calcium-sensing receptor regulates the NLRP3 inflammasome through Ca2+ and cAMP. Nature 2012, 492, 123–127. [Google Scholar] [CrossRef]
  139. Syrjanen, J.L.; Michalski, K.; Chou, T.-H.; Grant, T.; Rao, S.; Simorowski, N.; Tucker, S.J.; Grigorieff, N.; Furukawa, H. Structure and assembly of calcium homeostasis modulator proteins. Nat. Struct. Mol. Biol. 2020, 27, 150–159. [Google Scholar] [CrossRef]
  140. Dreses-Werringloer, U.; Lambert, J.-C.; Vingtdeux, V.; Zhao, H.; Vais, H.; Siebert, A.; Jain, A.; Koppel, J.; Rovelet-Lecrux, A.; Hannequin, D.; et al. A Polymorphism in CALHM1 Influences Ca2+ Homeostasis, Aβ Levels, and Alzheimer’s Disease Risk. Cell 2008, 133, 1149–1161. [Google Scholar] [CrossRef] [Green Version]
  141. Jun, M.; Xiaolong, Q.; Chaojuan, Y.; Ruiyuan, P.; Shukun, W.; Junbing, W.; Liao, Y.; Hong, C.; Jinbo, C.; Rong, W.; et al. Calhm2 governs astrocytic ATP releasing in the development of depression-like behaviors. Mol. Psychiatry 2017, 23, 883–891. [Google Scholar] [CrossRef] [PubMed]
  142. Cheng, J.; Dong, Y.; Ma, J.; Pan, R.; Liao, Y.; Kong, X.; Li, X.; Li, S.; Chen, P.; Wang, L.; et al. Microglial Calhm2 regulates neuroinflammation and contributes to Alzheimer’s disease pathology. Sci. Adv. 2021, 7. [Google Scholar] [CrossRef] [PubMed]
  143. McLarnon, J.G. Microglial Store-operated Calcium Signaling in Health and in Alzheimer’s Disease. Curr. Alzheimer Res. 2021, 17, 1057–1064. [Google Scholar] [CrossRef]
  144. Deming, Y.; Li, Z.; Benitez, B.A.; Cruchaga, C. Triggering receptor expressed on myeloid cells 2 (TREM2): A potential therapeutic target for Alzheimer disease? Expert Opin. Ther. Targets 2018, 22, 587–598. [Google Scholar] [CrossRef]
  145. Jairaman, A.; McQuade, A.; Granzotto, A.; Kang, Y.J.; Chadarevian, J.P.; Gandhi, S.; Parker, I.; Smith, I.; Cho, H.; Sensi, S.L.; et al. TREM2 regulates purinergic receptor-mediated calcium signaling and motility in human iPSC-derived microglia. eLife 2022, 11, e73021. [Google Scholar] [CrossRef] [PubMed]
  146. Vardjan, N.; Verkhratsky, A.; Zorec, R. Astrocytic Pathological Calcium Homeostasis and Impaired Vesicle Trafficking in Neurodegeneration. Int. J. Mol. Sci. 2017, 18, 358. [Google Scholar] [CrossRef] [PubMed]
  147. Grolla, A.A.; Fakhfouri, G.; Balzaretti, G.; Marcello, E.; Gardoni, F.; Canonico, P.L.; DiLuca, 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]
  148. O Jalonen, T.; Charniga, C.J.; Wielt, D.B. β-Amyloid peptide-induced morphological changes coincide with increased K+ and Cl− channel activity in rat cortical astrocytes. Brain Res. 1997, 746, 85–97. [Google Scholar] [CrossRef]
  149. Abramov, A.; Canevari, L.; Duchen, M. Changes in Intracellular Calcium and Glutathione in Astrocytes as the Primary Mechanism of Amyloid Neurotoxicity. J. Neurosci. 2003, 23, 5088–5095. [Google Scholar] [CrossRef]
  150. Alberdi, E.; Wyssenbach, A.; Alberdi, M.; Sánchez-Gómez, M.V.; Cavaliere, F.; Rodríguez, J.J.; Verkhratsky, A.; Matute, C. Ca2+-dependent endoplasmic reticulum stress correlates with astrogliosis in oligomeric amyloid β-treated astrocytes and in a model of Alzheimer’s disease. Aging Cell 2013, 12, 292–302. [Google Scholar] [CrossRef] [Green Version]
  151. 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]
  152. Ahmadpour, N.; Kantroo, M.; Stobart, J.L. Extracellular Calcium Influx Pathways in Astrocyte Calcium Microdomain Physiology. Biomolecules 2021, 11, 1467. [Google Scholar] [CrossRef]
  153. Cho, J.; Huh, Y. Astrocytic Calcium Dynamics Along the Pain Pathway. Front. Cell. Neurosci. 2020, 14, 594216. [Google Scholar] [CrossRef] [PubMed]
  154. Anekonda, T.S.; Quinn, J.; Harris, C.; Frahler, K.; Wadsworth, T.L.; Woltjer, R.L. L-type voltage-gated calcium channel blockade with isradipine as a therapeutic strategy for Alzheimer’s disease. Neurobiol. Dis. 2011, 41, 62–70. [Google Scholar] [CrossRef] [PubMed]
  155. Demuro, A.; Parker, I.; Stutzmann, G.E. Calcium Signaling and Amyloid Toxicity in Alzheimer Disease. J. Biol. Chem. 2010, 285, 12463–12468. [Google Scholar] [CrossRef]
  156. Arispe, N.; Rojas, E.; Pollard, H.B. Alzheimer disease amyloid beta protein forms calcium channels in bilayer membranes: Blockade by tromethamine and aluminum. Proc. Natl. Acad. Sci. USA 1993, 90, 567–571. [Google Scholar] [CrossRef]
  157. Guan, P.-P.; Cao, L.-L.; Wang, P. Elevating the Levels of Calcium Ions Exacerbate Alzheimer’s Disease via Inducing the Production and Aggregation of β-Amyloid Protein and Phosphorylated Tau. Int. J. Mol. Sci. 2021, 22, 5900. [Google Scholar] [CrossRef]
  158. Supnet, C.; Grant, J.; Kong, H.; Westaway, D.; Mayne, M. Amyloid-β-(1-42) Increases Ryanodine Receptor-3 Expression and Function in Neurons of TgCRND8 Mice. J. Biol. Chem. 2006, 281, 38440–38447. [Google Scholar] [CrossRef]
  159. Skowrońska, K.; Kozłowska, H.; Albrecht, J. Neuron-derived factors negatively modulate ryanodine receptor-mediated calcium release in cultured mouse astrocytes. Cell Calcium 2020, 92, 102304. [Google Scholar] [CrossRef]
  160. Oules, B.; Del Prete, D.; Greco, B.; Zhang, X.X.; Lauritzen, I.; Sevalle, J.; Moreno, S.; Paterlini-Bréchot, P.; Trebak, M.; Checler, F.; et al. Ryanodine Receptor Blockade Reduces Amyloid- Load and Memory Impairments in Tg2576 Mouse Model of Alzheimer Disease. J. Neurosci. 2012, 32, 11820–11834. [Google Scholar] [CrossRef]
  161. Briggs, C.A.; Chakroborty, S.; Stutzmann, G.E. Emerging pathways driving early synaptic pathology in Alzheimer’s disease. Biochem. Biophys. Res. Commun. 2016, 483, 988–997. [Google Scholar] [CrossRef] [PubMed]
  162. Del Prete, D.; Checler, F.; Chami, M. Ryanodine receptors: Physiological function and deregulation in Alzheimer disease. Mol. Neurodegener. 2014, 9, 21. [Google Scholar] [CrossRef] [PubMed]
  163. Liang, J.; Kulasiri, D.; Samarasinghe, S. Ca2+ dysregulation in the endoplasmic reticulum related to Alzheimer’s disease: A review on experimental progress and computational modeling. Biosystems 2015, 134, 1–15. [Google Scholar] [CrossRef]
  164. Minicucci, J.; Alfond, M.; Demuro, A.; Gerberry, D.; Latulippe, J. Quantifying the dose-dependent impact of intracellular amyloid beta in a mathematical model of calcium regulation in xenopus oocyte. PLoS ONE 2021, 16, e0246116. [Google Scholar] [CrossRef] [PubMed]
  165. Tong, C.K.B.; Wu, A.J.; Li, M.; Cheung, K.-H. Calcium signaling in Alzheimer’s disease & therapies. Biochim. Biophys. Acta (BBA) Bioenerg. 2018, 1865, 1745–1760. [Google Scholar] [CrossRef]
  166. DeMuro, A.; Parker, I. Cytotoxicity of Intracellular A 42 Amyloid Oligomers Involves Ca2+ Release from the Endoplasmic Reticulum by Stimulated Production of Inositol Trisphosphate. J. Neurosci. 2013, 33, 3824–3833. [Google Scholar] [CrossRef]
  167. Green, K.N.; LaFerla, F.M. Linking Calcium to Aβ and Alzheimer’s Disease. Neuron 2008, 59, 190–194. [Google Scholar] [CrossRef]
  168. Paula-Lima, A.C.; Adasme, T.; SanMartín, C.; Sebollela, A.; Hetz, C.; Carrasco, M.A.; Ferreira, S.T.; Hidalgo, C. Amyloid β-Peptide Oligomers Stimulate RyR-Mediated Ca2+ Release Inducing Mitochondrial Fragmentation in Hippocampal Neurons and Prevent RyR-Mediated Dendritic Spine Remodeling Produced by BDNF. Antioxid. Redox Signal. 2011, 14, 1209–1223. [Google Scholar] [CrossRef]
  169. Shtifman, A.; Ward, C.W.; Laver, D.R.; Bannister, M.L.; Lopez, J.R.; Kitazawa, M.; LaFerla, F.M.; Ikemoto, N.; Querfurth, H.W. Amyloid-β protein impairs Ca2+ release and contractility in skeletal muscle. Neurobiol. Aging 2010, 31, 2080–2090. [Google Scholar] [CrossRef]
  170. DeMuro, A.; Smith, M.; Parker, I. Single-channel Ca2+ imaging implicates Aβ1–42 amyloid pores in Alzheimer’s disease pathology. J. Cell Biol. 2011, 195, 515–524. [Google Scholar] [CrossRef] [Green Version]
  171. Ullah, G.; DeMuro, A.; Parker, I.; Pearson, J.E. Analyzing and Modeling the Kinetics of Amyloid Beta Pores Associated with Alzheimer’s Disease Pathology. PLoS ONE 2015, 10, e0137357. [Google Scholar] [CrossRef] [PubMed]
  172. Fani, G.; Mannini, B.; Vecchi, G.; Cascella, R.; Cecchi, C.; Dobson, C.M.; Vendruscolo, M.; Chiti, F. Aβ Oligomers Dysregulate Calcium Homeostasis by Mechanosensitive Activation of AMPA and NMDA Receptors. ACS Chem. Neurosci. 2021, 12, 766–781. [Google Scholar] [CrossRef] [PubMed]
  173. Pannuzzo, M. Beta-amyloid pore linked to controlled calcium influx into the cell: A new paradigm for Alzheimer’s Disease. Alzheimer’s Dement. 2021, 18, 191–196. [Google Scholar] [CrossRef] [PubMed]
  174. Latulippe, J.; Lotito, D.; Murby, D. A mathematical model for the effects of amyloid beta on intracellular calcium. PLoS ONE 2018, 13, e0202503. [Google Scholar] [CrossRef]
  175. 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]
  176. Chen, G.F.; Xu, T.H.; Yan, Y.; Zhou, Y.R.; Jiang, Y.; Melcher, K.; Xu, H.E. Amyloid beta: Structure, biology and structure-based therapeutic development. Acta Pharmacol. Sin. 2017, 38, 1205–1235. [Google Scholar] [CrossRef]
  177. Fontana, I.C.; Zimmer, A.R.; Rocha, A.S.; Gosmann, G.; Souza, D.O.; Lourenco, M.V.; Ferreira, S.T.; Zimmer, E.R. Amyloid-β oligomers in cellular models of Alzheimer’s disease. J. Neurochem. 2020, 155, 348–369. [Google Scholar] [CrossRef]
  178. Cascella, R.; Cecchi, C. Calcium Dyshomeostasis in Alzheimer’s Disease Pathogenesis. Int. J. Mol. Sci. 2021, 22, 4914. [Google Scholar] [CrossRef]
  179. Alberdi, E.; Sánchez-Gómez, M.V.; Cavaliere, F.; Pérez-Samartín, A.; Zugaza, J.L.; Trullas, R.; Domercq, M.; Matute, C. Amyloid β oligomers induce Ca2+ dysregulation and neuronal death through activation of ionotropic glutamate receptors. Cell Calcium 2010, 47, 264–272. [Google Scholar] [CrossRef]
  180. Wang, Y.; Shi, Y.; Wei, H. Calcium Dysregulation in Alzheimer’s Disease: A Target for New Drug Development. J. Alzheimer’s Dis. Park. 2017, 7, 374. [Google Scholar] [CrossRef]
  181. Ge, M.; Zhang, J.; Chen, S.; Huang, Y.; Chen, W.; He, L.; Zhang, Y. Role of Calcium Homeostasis in Alzheimer’s Disease. Neuropsychiatr. Dis. Treat. 2022, 18, 487–498. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Crosstalk between Ca2+ dysregulation and neuroinflammation in Alzheimer’s disease. Among many other possible causes, the accumulation of amyloid-β (Aβ) plaques and neurofibrillary tangles (NFTs) in the AD brain, besides inducing microglia activation, bring about Ca2+ dysregulation in astrocytes. Ca2+ dysregulation increase calcineurin (CN) activity in astrocytes with AD, where CN assumes control over gene regulation through the activation of nuclear factor of activated T cells (NFAT) transcription factors, leading to the production and release of several inflammatory mediators, such as cytokines. In turn, inflammatory factors induce Ca2+ dysregulation and activation of the MAP kinase signalling pathways (p38) in neurons, leading to increased CN activity. Hyperactivation of CN in neurons determines transcriptional repression of several genes, especially those related to synapse function and integrity, causing synaptic plasticity deficits, neurite degeneration, and cognitive impairment. These parameters further participate in the activation of nearby astrocytes leading to the dysregulation of astrocytic Ca2+ and hyperactivity of astrocytic CN that, through the regulation of the transcriptional induction of genes, sustain glial activation and neuroinflammation, negatively affecting neurons and exacerbating neuronal Ca2+ dysregulation. Perpetuation of this vicious cycle drives the progression of AD pathology.
Figure 1. Crosstalk between Ca2+ dysregulation and neuroinflammation in Alzheimer’s disease. Among many other possible causes, the accumulation of amyloid-β (Aβ) plaques and neurofibrillary tangles (NFTs) in the AD brain, besides inducing microglia activation, bring about Ca2+ dysregulation in astrocytes. Ca2+ dysregulation increase calcineurin (CN) activity in astrocytes with AD, where CN assumes control over gene regulation through the activation of nuclear factor of activated T cells (NFAT) transcription factors, leading to the production and release of several inflammatory mediators, such as cytokines. In turn, inflammatory factors induce Ca2+ dysregulation and activation of the MAP kinase signalling pathways (p38) in neurons, leading to increased CN activity. Hyperactivation of CN in neurons determines transcriptional repression of several genes, especially those related to synapse function and integrity, causing synaptic plasticity deficits, neurite degeneration, and cognitive impairment. These parameters further participate in the activation of nearby astrocytes leading to the dysregulation of astrocytic Ca2+ and hyperactivity of astrocytic CN that, through the regulation of the transcriptional induction of genes, sustain glial activation and neuroinflammation, negatively affecting neurons and exacerbating neuronal Ca2+ dysregulation. Perpetuation of this vicious cycle drives the progression of AD pathology.
Cells 11 02728 g001
Figure 2. Schematic representation of Ca2+ homeostasis dysregulation in microglia and astrocytes in AD. P2X7R, purinergic P2X7 receptor; NLRP3, NOD-like receptor pyrin domain containing 3; Calhm1, Ca2+ homeostasis modulator family member 1; TREM2, triggering receptor expressed on myeloid cells 2; pCREB, phosphorylated cAMP response element binding protein; Calhm2, Ca2+ homeostasis modulator family member 2; mGluR5, metabotropic glutamate receptor 5; IP3R1, Inositol 1,4,5-trisphosphate (IP3) receptor type1; L-type VGGCs, L-type voltage-gated Ca2+ channels; RyRs, Ryanodine receptors. Created with BioRender.com; accessed on 29 July 2022.
Figure 2. Schematic representation of Ca2+ homeostasis dysregulation in microglia and astrocytes in AD. P2X7R, purinergic P2X7 receptor; NLRP3, NOD-like receptor pyrin domain containing 3; Calhm1, Ca2+ homeostasis modulator family member 1; TREM2, triggering receptor expressed on myeloid cells 2; pCREB, phosphorylated cAMP response element binding protein; Calhm2, Ca2+ homeostasis modulator family member 2; mGluR5, metabotropic glutamate receptor 5; IP3R1, Inositol 1,4,5-trisphosphate (IP3) receptor type1; L-type VGGCs, L-type voltage-gated Ca2+ channels; RyRs, Ryanodine receptors. Created with BioRender.com; accessed on 29 July 2022.
Cells 11 02728 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Di Benedetto, G.; Burgaletto, C.; Bellanca, C.M.; Munafò, A.; Bernardini, R.; Cantarella, G. Role of Microglia and Astrocytes in Alzheimer’s Disease: From Neuroinflammation to Ca2+ Homeostasis Dysregulation. Cells 2022, 11, 2728. https://doi.org/10.3390/cells11172728

AMA Style

Di Benedetto G, Burgaletto C, Bellanca CM, Munafò A, Bernardini R, Cantarella G. Role of Microglia and Astrocytes in Alzheimer’s Disease: From Neuroinflammation to Ca2+ Homeostasis Dysregulation. Cells. 2022; 11(17):2728. https://doi.org/10.3390/cells11172728

Chicago/Turabian Style

Di Benedetto, Giulia, Chiara Burgaletto, Carlo Maria Bellanca, Antonio Munafò, Renato Bernardini, and Giuseppina Cantarella. 2022. "Role of Microglia and Astrocytes in Alzheimer’s Disease: From Neuroinflammation to Ca2+ Homeostasis Dysregulation" Cells 11, no. 17: 2728. https://doi.org/10.3390/cells11172728

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