Next Article in Journal
Desmin and Plectin Recruitment to the Nucleus and Nuclei Orientation Are Lost in Emery-Dreifuss Muscular Dystrophy Myoblasts Subjected to Mechanical Stimulation
Previous Article in Journal
In Vivo and In Vitro Pro-Fibrotic Response of Lung-Resident Mesenchymal Stem Cells from Patients with Idiopathic Pulmonary Fibrosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulation of Microglial Function by ATP-Gated P2X7 Receptors: Studies in Rat, Mice and Human

by
Manju Tewari
*,†,
Stephanie Michalski
and
Terrance M. Egan
Department of Pharmacology and Physiology, and The Henry and Amelia Nasrallah Institute for Translational Neuroscience, Saint Louis University School of Medicine, St. Louis, MO 63104, USA
*
Author to whom correspondence should be addressed.
Present address: Department of Biotechnology, Kumaun University, Bhimtal Campus, Bhimtal 263136, Uttarakhand, India.
Cells 2024, 13(2), 161; https://doi.org/10.3390/cells13020161
Submission received: 8 November 2023 / Revised: 10 January 2024 / Accepted: 12 January 2024 / Published: 16 January 2024

Abstract

:
P2X receptors are a family of seven ATP-gated ion channels that trigger physiological and pathophysiological responses in a variety of cells. Five of the family members are sensitive to low concentrations of extracellular ATP, while the P2X6 receptor has an unknown affinity. The last subtype, the P2X7 receptor, is unique in requiring millimolar concentrations to fully activate in humans. This low sensitivity imparts the agonist with the ability to act as a damage-associated molecular pattern that triggers the innate immune response in response to the elevated levels of extracellular ATP that accompany inflammation and tissue damage. In this review, we focus on microglia because they are the primary immune cells of the central nervous system, and they activate in response to ATP or its synthetic analog, BzATP. We start by introducing purinergic receptors and then briefly consider the roles that microglia play in neurodevelopment and disease by referencing both original works and relevant reviews. Next, we move to the role of extracellular ATP and P2X receptors in initiating and/or modulating innate immunity in the central nervous system. While most of the data that we review involve work on mice and rats, we highlight human studies of P2X7R whenever possible.

1. Introduction

Nucleotide triphosphates play many roles in the everyday physiology of all cells. They are vital components of the genetic code, essential cofactors for countless enzymatic reactions, and fundamental phosphate donors in biosynthesis [1]. Adenosine triphosphate (ATP) is particularly important because it functions both inside and outside the cell. Intracellular ATP provides the power needed to drive a vast array of energetically unfavorable chemical reactions within cells, whereas extracellular ATP (eATP) and its metabolites activate plasmalemmal purinergic receptors that initiate and modulate cellular function [2,3]. The purinergic receptor superfamily contains three branches that differ in structure, function, and pharmacology [4,5]. The first class, called P1 receptors (more commonly called “A receptors”), are metabotropic receptors activated by nanomolar to low micromolar concentrations of extracellular adenosine, a common by-product of dephosphorylation of eATP. Four subtypes (A1, A2A, A2B, and A3 receptors) are coupled to G-proteins, and microglia express all four [6,7,8,9,10,11,12]. The second class of purinergic receptors, called P2Y receptors (P2YRs), resemble adenosine receptors in signaling through the activation of G-protein coupled receptors. This subfamily is composed of eight mammalian subtypes (P2Y1R, P2Y2R, P2Y4R, P2Y6R, P2Y11R, P2Y12R, P2Y13R, and P2Y14R), many of which are expressed in microglia [13,14,15]. Individual family members have a preferred native agonist of ADP, ATP, UDP, or UTP [16]. P2YRs play essential roles in microglial chemotaxis, phagocytosis, and the release of pro-and anti-inflammatory cytokines [17,18,19,20].

P2X Receptors

P2X receptors (P2XRs) form the third class of purinergic receptors. They are a family of seven ATP-gated ion channels, designated P2X1R–P2X7R, that depolarize cells by increasing membrane permeability to Na+, K+, and Ca2+ [21]. In humans, five of the seven subtypes form homotrimeric complexes capable of stimulating transmembrane current in response to eATP. Human P2X5R and P2X6R are the exceptions. In the case of human P2X5R, a single-nucleotide polymorphism in p2rx5 prevents transcription of the exon 10 resulting in an incomplete protein that lacks a part of the pore-lining second transmembrane domain (TM2); the resulting product properly traffics to the surface membrane but fails to respond to eATP [22]. When engineered to contain all 13 exons, the transfection of cDNA encoding the full-length human P2X5R results in the expression of functional eATP-gated receptor which, unlike other family members, displays significant Cl permeability (i.e., relative Cl to Na+ permeability equals 0.5 for human P2X5R as compared to <0.1 for other P2XR subtypes) [23]. While the expression of the functional, full-length P2X5R does not normally occur in humans, a single G > T polymorphism would make it possible, an outcome worth considering when studying the effects of eATP on human immune cells that show abundant levels of P2X5R mRNA [22].
Like the case of human P2X5Rs, the expression of the gene encoding the human P2X6R also does not result in a functional receptor. The lack of ATP-gated current reflects the fact that monomeric P2X6 subunits are retained in the endoplasmic reticulum and require a partner (either P2X1, P2X2, or P2X5) to shuttle it to the plasma membrane as a heteromeric receptor [24,25]. Indeed, the ability to form heteromeric receptors is a common feature of most P2XRs [26], including those found in native tissues [27]. The P2X7 receptor is the only receptor for which this is not true. It is incapable of pairing with other P2XR subtypes and always expresses as homotrimeric P2X7R [26,28]. There are 11 P2X7R rodent splice variants (P2X7aR–P2X7kR) and eight human variants (P2X7AR–P2X7HR), of which only P2X7AR and P2X7BR form functional complexes gated by eATP in humans [29]. P2X7BR differs from P2X7AR in four ways: it shows higher mRNA transcript numbers; it has a truncated C-terminal cytoplasmic domain; it does not induce membrane permeabilization (see below); and, it is less sensitive to eATP [30]. Co-expression of P2X7AR and P2X7BR produces a leftward shift in the eATP concentration–membrane current response curve and increased uptake of fluorescent dyes, demonstrating that the co-assembly of the two splice variants into a single complex results in a heteromeric P2X7A/P2X7B receptor with enhanced function [31].
P2X7ARs are unique in other ways too. Chief amongst these differences is a low sensitivity to eATP by comparison to other family members, absence of desensitization, and persistent current facilitation in response to sustained agonist ligation [32]. Like all functional P2XRs, the P2X7aR is composed of three subunits, with each subunit spanning the membrane twice. The N- and C-termini are intracellular, and the orthosteric eATP-binding site is formed at the extracellular interface of adjoining subunits [33,34]. The C-terminus is ~200 amino acids longer than those of P2X1R–P2X6R and has sites that bind accessory proteins, modulate receptor function via post-translational modification, regulate desensitization, and help target the receptor to the surface membrane [30,35,36,37]. Recent structural data obtained from single-particle cryo-electron microscopy were the first to describe the complete structure of a full-length wild-type P2X7R with intact cytoplasmic domains, and this landmark paper from the Mansoor laboratory provides plausible molecular explanations for two of the unique properties mentioned above [34]. First, unlike other P2XRs, the long C-terminus of the P2X7aR contains a unique 18-amino acid cysteine-rich domain that the authors call the “C-cys anchor”, which acts as a molecular hinge that prevents desensitization by using palmitoyl groups to anchor the pore-lining TM2 and the cytoplasmic domains to the inner leaflet of the lipid membrane. This lack of desensitization imparts P2X7aRs with the profile of sustained application that is a hallmark of P2X7aR’s pharmacological fingerprint. Second, P2X7aRs show an unusually low affinity for eATP, which traditionally has been explained by assuming that fully ionized ATP (i.e., ATP4−), a form of the nucleotide that is in low abundance in the presence of the millimolar concentrations of divalent cations (predominately Ca2+ and Mg2+) normally present in interstitial fluids, is the natural agonist [34,38]. Now, the solved structure offers another possibility: the entrance to the extracellular ligand binding pocket of the P2X7 receptor is narrow in comparison to other subtypes, suggesting that the low affinity of eATP for P2X7aRs might result, in part, from limited access to its binding site [34]. This hypothesis could be directly tested by comparing the atomic scale structures of the P2X7aR to the “k” splice variant of murine P2X7Rs (i.e., murine P2X7kR) that show a remarkable eight-fold difference in agonist sensitivity [39].
Human tissues do not express the more sensitive “K” splice variant, and therefore, relatively high concentrations of eATP are needed to trigger a P2X7AR-mediated response in humans. The low sensitivity of the P2X7AR (the main focus of this review and henceforth simply referred to as “P2X7R”) is physiologically significant because it provides a means by which eATP can act as a danger-associated molecular pattern (DAMP) that signals cell damage. The extracellular concentration of eATP is low (~10 nM) in healthy tissue because of limited release and rapid hydrolysis by ectonucleotidases [40,41,42]. However, it rises when cytoplasmic ATP (~1–10 mM) is released through canonical (exocytosis) and non-canonical (cell lysis and passive transport through ion channels) pathways [43,44]. Any condition that permeabilizes membranes results in the outward leak of intracellular ATP [42,45]. The result is an increase in eATP at sites of inflammation and tissue damage [46] that, for example, can reach concentrations as high as hundreds of micromolar in the microenvironment of tumors [47,48,49]. In the CNS, multiple pathologies, including ischemia, trauma, oxidative stress, hypoxia, neurodegenerative insult, cellular stress, and epilepsy, significantly elevate concentrations of eATP and its metabolites in the interstitial fluid [50]. Low concentrations of eATP (tens of micromolar) act as short-range “find me” signals directing motile phagocytes to damaged tissue by activating metabotropic P2YRs [51,52,53,54,55]. Higher concentrations (hundreds of micromolar) activate P2X7Rs, which in some cases result in the translocation of phosphatidylserine, bestowing an “eat me” signal in the outer leaflet of the plasma membrane that targets cells for engulfment and cell death [56,57,58]. Interestingly, the activation of P2X7Rs does not trigger phosphatidylserine translocation or phagocytosis in cultured human microglia. Instead, eATP decreases the uptake of fluorescently labeled E. coli bioparticles [59], which supports the hypothesis that the closed/inactive conformation of the P2X7R acts as a scavenger receptor for innate phagocytosis in the absence of eATP [60].
One final feature sets P2X7Rs apart from other members of the P2XR family. Some cells, particularly those associated with the immune system, express a P2X7R that does not initiate a measurable response to the application of eATP [61,62]. nfP2X7R (“nf” for non-functional) is normally retained intracellularly and released to the membrane upon stimulation with concentrations of eATP > 0.5 mM. It is expressed in a wide range of cancers where it promotes tumor cell survival [62,63]. Relatively little is known about how this unique receptor works, so further work is warranted. To date, there are no reports of expression of nfP2X7R in microglia of any species.

2. Microglia

The human brain contains ~160 billion cells. Roughly half of these are neurons, with the remainder comprised of endothelium (~20 billion) and neuroglia [64]. Originally, neuroglia were thought to simply provide physical and nutritional support to neurons [65]. However, in light of an expanding body of evidence, it is now abundantly clear that these cells play substantive roles in neural development, plasticity, homeostasis, and disease [20,66,67].
Approximately 5–20% of the ~60 billion glial cells in the human brain are microglia [68]. Microglia begin life as erythromyeloid progenitor-derived yolk sac macrophages that populate the nervous system during early embryogenesis [67,69]; the idea that fetal-liver-derived monocytes also contribute is postulated but controversial [70]. In humans, migration begins at ~4.5 gestational weeks, with a second wave occurring ~8 weeks later [71,72]. The precise route by which they enter the CNS is unknown but may involve the meninges, ventricles, and vasculature [67,73]. Microglia colonize the CNS in a grid-like fashion and quickly develop phenotypes that vary as a function of the soluble components of CNS microenvironments [74,75]. These environmental signals are sensed by a family of microglial proteins collectively known as the “sensome”, which includes receptors for pattern recognition signals, danger-associated molecular patterns (DAMPs), Fc fragments, chemokines, cytokines, and extracellular matrix proteins [76]. Purinergic receptors, including the P2X7R, form a significant portion (~8%) of the sensome, highlighting the importance of eATP and its metabolites in directing microglial function. Regional differences in postnatal microglial are reported in rodents [77,78,79,80,81] and humans [82,83,84] but remain controversial [67,75]. For example, the ability of eATP stimulation of microglia to kill co-cultured neurons depends in part on the section of the brain from which the microglia were harvested [77]. In addition, microglia show developmental differences in regional densities, transcriptomes, activation states, and functional properties that vary between males and females [85,86,87,88], including humans [82,89].
In postnatal infants and healthy adult animals, an intact blood–brain barrier prevents further invasion by components of the peripheral immune system, and microglia depend on proliferation and apoptosis to sustain a relatively stable population of mature cells [90,91,92]. Absolute rates of proliferation are controversial [93] but appear to vary by brain region [94]. The average lifespan is thought to be ~4.2 years in humans, with about 2% of the total microglial population proliferating at any one time [91,95]. It is worth noting that many diseases compromise the blood–brain barrier; glioblastoma is an example. In these cases, peripheral blood monocytes and non-parenchymal macrophages migrate into the CNS, where they differentiate into cells that can be hard to distinguish from resident microglia [74,75,96,97,98]. Further, serum permeates the compromised blood–brain barrier, exposing the CNS parenchyma to regulatory signals that it normally would not encounter [99]. Because it promotes cell division [100], the leak of serum into the CNS environment might explain the increased microglial proliferation and phagocytosis observed following stroke and trauma [100].
Microglia express a common pool of developmental and homeostatic genes in a wide range of species, including humans. However, unlike other animals, human microglia show significant heterogeneity with clearly delineated subsets that are unrelated to sex [101]. Further, humans express a greater number of genes conferring susceptibility to Alzheimer’s and Parkinson’s diseases than mice and rats (see below). Thus, although procurement of tissue samples of the human brain is problematic and the range of possible in situ experimental manipulations is severely limited by comparison to experiments on rodents, more data from work on human tissue is needed to better understand how microglia contribute to human CNS disease [102].

2.1. Microglia Express Multiple Subtypes of Purinergic Receptors

Microglia express key elements of the “purinome” [103] and are capable of releasing, metabolizing, and sensing ATP [15,104]. They express a number of ectonucleotidases, including CD39, which metabolizes ATP and ADP to AMP; NPP1, which metabolizes ATP to AMP; and CD73, which metabolizes AMP to adenosine [105,106]. Microglia also express relatively high mRNA levels for metabotropic (adora2A, adora3, p2ry2, p2ry6, p2ry12, and p2ry13) and ionotropic (p2rx4 and p2rx7) purinergic receptors [15,107,108]. The additional receptors identified by functional studies include A1, A2A, P2X1, P2Y1 and P2Y14 receptors [109,110,111,112]. The relative abundance of each receptor subtype depends in part on the sex of the animal [86] and the activation state of the microglia [110]. For example, seizures in a rodent model of epilepsy cause the upregulation of mRNA expression for P2Y6Rs, P2Y12Rs, P2Y13Rs, P2X4Rs, and P2X7Rs in hippocampal microglia [113].
It is difficult to assign distinct physiological responses to specific purinergic receptors. However, some generalizations hold true. Adenosinergic A3Rs regulate microglial process extension and migration of microglia [11,114]. Metabotropic P2Y6Rs underlie phagocytosis [115], and P2Y12Rs mediate cell migration [15,52]. Ionotropic P2X4Rs [116,117] and P2X7Rs [118,119,120] are linked to neuropathic pain. Additionally, P2X7Rs mediate the release of proinflammatory cytokines in response to tissue damage and inflammation [108,121,122].

2.2. Healthy Prenatal CNS

Microglia, under the direction of the chemotactic receptor CX3CR1 [69,123,124], infiltrate the CNS as neuronal circuits begin to assemble [71,125] and play essential roles in the development of the nervous system [126,127,128]. In rats and monkeys, they selectively colonize proliferative zones during the late stages of cortical neurogenesis and limit the number of neurons by phagocytosing precursor cells [129]. Further, they influence the differentiation of both astrocytes and oligodendrocytes in mice [72,130] and humans [72] and may be angiogenic [131,132]. Colonization is dimorphic, with sex-dependent differences in regional densities that vary with time [133,134]. For example, male and female rodents have equivalent densities of activated microglia in the amygdala, hippocampus, and parietal cortex at birth. However, density is higher in males in comparison to females at postnatal day 4, and higher in females than males in mature adults [85]. The chemotactic signals that attract microglia to the brain are incompletely characterized but certainly include CX3CL1 (also known as fractalkine). Genetically modified mice lacking the fractalkine/CX3CR1 receptor show a significant reduction in the number of microglia in the brain and a concurrent increase in the density of dendritic spines, suggesting that CX3CR1 guides the migration of the microglia responsible for pruning synapses in early development [123,124,135]. Nucleotides also play a role, as the high concentration of extracellular nucleotides that accompanies developmental neuronal apoptosis is suggested to act as a chemotactic signal for recruitment of microglial precursors into zebrafish brain [136]. The identity of the responsible nucleotide is unknown but probably derived from ATP, which acts through metabotropic P2Y12Rs to direct microglia to sites of CNS injury in adult murine brains [52,137].

2.3. Healthy Postnatal CNS

Immature microglia invade the CNS as irregular-shaped cells with blunt processes, a morphological profile that resembles activated microglia in adults [89,138,139,140]. They disburse in a uniform fashion, with typical cell-to-cell distances of about 50–60 µm in adult mice [141] and remain in place in the absence of a migratory signal or pathophysiological insult [142]. Their amoeboid shape changes by postnatal day 10 to a ramified morphology of radial extensions from a small rod-shaped cell soma, with a full extension completed by postnatal day 28 [73]. The first-order branches, which can reach a length of 50 µm, are highly motile, extending and retracting at a rate of 1.5 µm/min and capable of surveying the entire brain environment once every few hours [141,143]. Normal process extension and baseline surveillance depends on tonic activation of the two-pore K+ channel, THIK-1, which is responsible for setting the resting membrane potential in mouse microglia [19]. Thinner filopodia spread from the tips of the larger branches and extend and retract in a manner that is independent of THIK-1 but modulated by cAMP, enabling precise surveillance of the immediate environment under spatiotemporal control of intracellular cAMP microdomains [144]. Purinergic receptors also come into play in human brain slices, as low doses of ADP trigger process extension through the activation of P2Y12Rs and higher doses trigger process retraction through the activation of P2Y1Rs and P2Y13Rs [112]. Adenosine, a by-product of ATP metabolism, also triggers process retraction by activating Gs-protein-coupled A2A receptors [8].
Although microglia contact a variety of cell types, they preferentially target active neurons [145,146,147]. Approximately 90% of neocortical neurons in mice and humans are in contact with a microglial process at any one time, with typical lifespans of 7.5 and 25 min for dendritic and somatic connections, respectively [148]. Interestingly, somatic connections display a unique molecular architecture of apposing clusters of neuronal Kv2.1 K+ channels and microglial P2Y12 receptors, a pattern not seen in dendritic contacts. Neuronal mitochondria, vesicular nucleotide transporter (vNUT), and an ectonucleotidase (NTPDase1) are positioned close to the Kv2.1 clusters, suggesting that neurons signal increases in activity by triggering the vesicular release of mitochondrial-derived ATP, which is rapidly converted to ADP, the natural agonist for the P2Y12R [148].
As the primary immune cell of the CNS, microglia attack infections of the brain and spinal cord (see below). However, in the absence of infection, microglia remain active and play critical roles in synapse maturation and remodeling [149]. Microglia express and release neuroprotective chemicals that promote the development of glial precursor cells, neurons, and oligodendrocytes [138]. At the same time, neurons upregulate the expression of fractalkine at peak times of synapse maturation (P15 in mice), providing a strong signal for the migration of microglia to developing synapses [150]. Once in place, microglia engulf and phagocytose dendritic spines to reduce weaker afferent input and fine-tune synaptic connectivity [135]. In mice, immature synapses formed of inactive afferent fibers are opsonized by complement proteins C1q and C3 [151,152], selectively targeting them for engulfment by resident microglia [146,149,153,154]. Phagocytosis of synapses continues into adulthood, as healthy adults show selective elimination of newly formed synapses during REM sleep [155,156,157], which might be explained by the suppressive effects of the circadian release of noradrenaline and glucocorticoids on microglial surveillance during waking hours [158]. Further, recent work suggests that complement-dependent synapse elimination by microglia leads to the forgetting of previously learned contextual fear memory in adult mice [159]. At the same time, microglia strengthen active synapses by secreting neurotrophic factors such as insulin-like growth factor, nerve growth factor and brain-derived neurotrophic factor [142,160,161], promote the synchronized discharge of neighboring neurons [162], and play a critical role in learning and memory [163,164].

2.4. Infection and Disease

Microglia are the first line of defense against bacterial, fungal, and viral infections of the CNS [165]. They respond to the detection of highly conserved microbial motifs, known as pathogen-associated molecular patterns (PAMPs), with a change in morphology and up-regulation of defensive signaling pathways. Common PAMPs include components of bacterial cell walls such as lipopolysaccharides (LPS) from Gram-negative bacteria, lipoteichoic acid from Gram-positive bacteria, and double-stranded RNA from viruses [166]. PAMPs are recognized by Toll-like receptors (TLRs), a family of pattern recognition receptors (PRRs) found in the cytosol and membranes of immune cells [167], including the human microglia [168,169]. Specific PAMPs are recognized by specific PRRs, with bacterial lipoteichoic acid and fungal zymosan recognized by TLR2, viral dsRNA by TLR3, bacterial LPS by TLR4, and bacterial flagellin by TLR5 [170,171]. The activation of TLRs promotes NF-κB-dependent signaling cascades, resulting in the production of chemokines (MIP-2, MCP-1), proinflammatory cytokines (TNF-α, IL-1β), and reactive oxygen/nitrogen species [170]. In addition, infection begets inflammation, triggering the release of DAMPs such as eATP from aggravated tissues, which promote the migration of phagocytes to the site of injury. The final outcome is the elimination of the insult through programmed cell death and phagocytosis [70,167,172]. Interestingly, new evidence from engraphment studies suggests that reactive microglia retain the ability to revert to the homeostatic ramified state when placed in a permissive environment lacking activation signals, suggesting that microglia are capable of surviving infection and reestablishing active surveillance upon the removal of the offending PAMP/DAMP [74].
Microglia express genes tied to neurodegeneration [173], and microglial expansion and activation are common features of neurodegenerative diseases [174,175,176,177]. Amongst these cells is a subset originally identified in rodent models of Alzheimer’s disease, amyotrophic lateral sclerosis, multiple sclerosis, and aging [178] but present in humans too [179,180]. Called disease-associated microglia (DAMs), they show downregulation of homeostatic signature genes such as cx3cr1, purinergic p2ry12, and tmem119, and upregulation of genes associated with lysosomal, phagocytic, interferon response, and lipid metabolism pathways [70]. The conversion of resting ramified microglia to activated amoeboid DAMs involves the recognition of neurodegeneration-associated molecular patterns (NAMPs) present on damaged and dying neurons, extracellular protein aggregates, and products of lipid degradation [178]. The signaling pathways responsible for the change in phenotype are under investigation, with good evidence supporting a role for the Triggering Receptor Expressed on Myeloid Cells 2 (TREM2) in mediating increases in proliferation and phagocytosis that define reactive microgliosis [181]. For example, genetically modified rats lacking TREM2 fail to show the DAM phenotype [182]. Further, TREM2 binds oligomers of β amyloid, leading to the colocalization of DAMs and plaques in a rodent model of Alzheimer’s disease [183,184,185,186]. Colocalization provides a protective shield that prevents neurons from exposure and positions microglia close to fibrillar plaques to facilitate phagocytosis [175]. Lastly, microglia express and release Apolipoprotein E (ApoE), a putative TREM2 agonist, and polymorphisms in the Apoe gene are a major risk determinate of Alzheimer’s disease [187,188]. Apoe is upregulated in DAMs, suggesting that the agonist, ApoE, and its receptor, TREM2, form a positive feedback loop that facilitates transition to the reactive DAM phenotype [175,189].

3. P2X7R and Microglia

Microglia show constitutive expression of p2rx4 and p2rx7 in both rodents and humans [20,108]. Microglial P2X4Rs play critical roles in rodent models of neuropathic pain [117,190], experimental allergic encephalomyelitis [191,192], and chronic migraine [193]. In keeping with these studies, relatively low concentrations of eATP (≤100 µM) evoke desensitizing non-selective inward currents and sustained outward K+ currents in murine microglia that most likely result from activation of P2X4Rs and P2Y12Rs, respectively [18,19,194,195,196,197]. Surprisingly, these same concentrations of eATP fail to trigger inward membrane current or Ca2+ influx in cultured human microglia, suggesting that either the p2rx4 is downregulated in culture, the mRNA is not translated, or the properly translated protein is not trafficked to the cell surface membrane [59,198].
In contrast, and as expected from work on mice [199], short applications of higher concentrations of eATP (≥1 mM) or the higher affinity ATP analog, 2′,3′-O-(benzoyl-4benzoyl)-ATP (BzATP), evoke inward currents in human microglia with properties expected of a P2X7R-mediated response (Figure 1); the current is cation non-selective, carried in part by Ca2+, facilitated during prolonged exposure to agonists, and blocked by P2X7R antagonists [59]. The resting membrane potential of human microglia is unknown. In rodents, it varies with age [200] but averages around −40 mV [19,201]. At this potential, eATP activation of P2X7Rs causes Na+ and Ca2+ to rush into the cell as K+ exits. The net result is membrane depolarization. In rodents, the inward Ca2+ current increases the concentration of free intracellular Ca2+ ([Ca2+]i), which triggers cell cycle progression [202], the release of TNF-α [203] and plasminogen [204], activation of the transcription factor NFAT [205], disruption of the cytoskeleton [206], and the production of H2O2 [207]. At the same time, the outward K+ current decreases the concentration of intracellular K+ ([K+]i), leading to activation of the NLRP3 inflammasome and maturation and release of the proinflammatory cytokines, IL-1β and IL-18 [121]. While it is well accepted that the inflammasome activates when [K+]i drops below 90 mM [208], more recent evidence suggests that the P2X7R is not the primary K+ efflux pathway [209]. Rather, the two-pore K+ channels, THIK-1 and TWIK-2, are responsible in microglia [19,210,211] and macrophages [212], respectively. The data supporting a role for THIK-1 and TWIK-2 are convincing, and the conclusions are firm. However, it is unclear why K+ efflux through the P2X7R is not sufficient by itself. In the microglial study of Madry et al., the simple fact that 2 mM ATP did not evoke a P2X7R-mediated current is enough to eliminate this receptor from consideration [19]. The lack of response is surprising because others report robust eATP-gated currents with properties unique to P2X7Rs in murine and human microglia [59,199,213,214], perhaps suggesting that P2X7R expression resembles P2X4R expression in its sensitivity to the choice of animal, the activation state of the microglia, and/or the experimental protocols used to study the cell. Regardless, when present, activation of P2X7Rs results in a large efflux of 86Rb+, a proxy for K+, in J774 macrophages and presumably in microglia [215]. Further, eATP promotes recruitment and colocalization of microglial P2X7Rs and NLRP3 to discrete sites of the subplasmalemmal cytoplasm, suggesting that the inflammasome is positioned close enough to directly sense the P2X7R-mediated local drop in [K+]i [216].

3.1. Membrane Permeabilization and Cell Lysis

Applications of eATP that last longer than a few seconds result in membrane permeabilization, a hallmark property of P2X7R activation [217,218,219]. Permeabilization is the process by which eATP triggers membrane transport of hydrophilic solutes with molecular masses of <900 Da in a direction determined by their electrochemical potential. The process is reversible [32] and does not necessarily lead to cell death. The ability of eATP to permeabilize membranes was first discovered in mouse 3T3 fibroblasts [220], rat mast cells [221], and mouse J774 macrophages [215,222,223], and later described in mouse microglia [224]. Originally thought to be a unique property of P2X7Rs, it was subsequently shown to accompany the activation of purinergic P2X2Rs, P2X3Rs and P2X4Rs [225,226,227] as well as proton-gated TRPV1 receptors [228]. While the physiological consequence of membrane permeabilization is the loss of cytoplasmic components such as ATP, cGMP, glutamate, and spermidine, the underlying process is typically measured as uptake of polyatomic fluorescent biomarkers such as the nucleic acid stains ethidium bromide and YO-PRO-1 [218,219]. For example, in HEK293 cells expressing recombinant P2X7Rs, BzATP triggers an increase in fluorescence that develops slowly over the course of seconds to minutes as dye enters the cell and intercalates DNA [217]. The primary route of entry is unclear [218]. That some of the dye transits the P2X7R pore is convincingly proven using reconstituted receptors and liposomes that lack other proteins [229]. At the same time, eATP and P2X7Rs may activate secondary transport pathways. For example, pannexin-1 is a plasma membrane cation-selective channel related to gap junction proteins that shows a close membrane association with P2X7Rs. Pharmacologically inhibiting pannexin-1 blocks the initial phase of eATP-gated ethidium uptake in human lung alveolar macrophages, demonstrating that it is partially responsible for membrane permeabilization to cationic macromolecules in these cells [230]. The same drugs also block P2X7R-dependent release of IL-1β, suggesting that pannexin-1 is a critical component of eATP-mediated NLRP3 inflammasome activation [230,231]. In contrast, blocking pannexin-1 has no effect on eATP-mediated membrane permeabilization of monocyte-derived human macrophages [232] or cultured human microglia [59], suggesting that local paracrine signaling active in tissue microenvironments may recruit distinct permeabilization pathways in a tissue-specific manner. The hypothesis that eATP gates multiple transport portals is further supported by the finding that, in some cases, P2X7Rs increase membrane permeability to both cations (YO-PRO-1 and ethidium) and anions (glutamate and Lucifer yellow) in the same cell [223,233,234,235]. It is unlikely that large anions travel through P2X7Rs because these channels show a strong preference for cations [217]. In addition, the uptake of cations is temperature-dependent, whereas the uptake of anions is not [236], again supporting the presence of multiple permeation pathways in some cells. With specific regard to human microglia, eATP does not trigger uptake of large anions in cultured cells [59], although this may reflect the downregulation of genes encoding the necessary pathway; additional experiments on cells that more closely resemble the natural phenotype of in situ microglia are needed.
The physiological and pathophysiological outcomes of membrane permeabilization are uncertain [218] and have not been extensively investigated in microglia. As mentioned above, the process is reversible and does not necessarily lead to cell death. For example, a 30 min application of ATP or BzATP causes significant uptake of cationic YO-PRO-1 in human microglia without inducing the release of lactate dehydrogenase, demonstrating that permeabilization is not lytic under these conditions [59]. In contrast, longer incubations result in necrosis or apoptosis in murine microglia, a result that is prevented by pre-incubation of a P2X7R antagonist [43,237]. Cell lysis is blocked by P2X7R antagonists and absent in cells isolated from P2X7R−/− animals, suggesting that the activation of P2X7R is a critical component of the lytic pathway [108]. Interestingly, eATP-activation of P2X2Rs permeabilizes membranes but does not kill cells, suggesting that permeabilization and lysis proceed through different intracellular signaling pathways [226]. If cell death is not the ultimate consequence, then what purpose does permeabilization serve? The recent work of the Grutter laboratory suggests one possibility [227]. Spermidine is a naturally occurring intracellular polyamine that acts at extracellular sites to allosterically modulate ion channel gating [238]. To be effective, it must be secreted. Spermidine permeates multiple subtypes of P2XRs, including P2X7Rs, suggesting that these receptors represent an eATP-dependent egress pathway for polyamines, an effect that may help to explain the ability of P2X7Rs to modulate the activity of neighboring ion channels [219].

3.2. Membrane Blebbing and Microvesicular Shedding

Non-apoptotic membrane blebbing occurs when the actin cytoskeleton separates from the plasma membrane; detachment allows the hydrostatic pressure within the cell to push the membrane through the actin cortex, forming an outwardly facing membrane extrusion [239,240]. eATP, working through P2X7Rs, is a potent initiator of non-apoptotic membrane blebbing in many cell types, including human macrophages [232] and murine microglia [13,241,242]. Here, blebbing occurs within minutes of P2X7R activation [241] and reverses when the agonist is removed [232]. As is the case of membrane permeabilization (see above), the physiological sequela of membrane blebbing is uncertain. In tumor cells, blebs facilitate cytokinesis [243], and in streptolysin-permeabilized human embryonic kidney cells, they trap damaged membrane segments and limit further cellular damage [244]. In THP-1 monocytes [241] and primary mouse microglia [13], bleb formation provides a vehicle for IL-1β release. Here, eATP activation of P2X7Rs results in rapid movement of phosphatidylserine and acid sphingomyelinase to the outer leaflet of the plasma membrane, resulting in the formation of small (40–80 nm) membrane-derived microvesicles that contain IL-1β. Surprisingly, eATP does not cause blebbing in cultured human microglia [59].

3.3. Cytokines and Reactive Oxygen Species (ROS)

The ability of eATP acting on purinergic P2 receptors to function as a DAMP has long established a role for eATP in regulating neuroinflammatory immune responses [167,245]. The immune response is characterized by a proinflammatory state, which is driven by immune cell activation and the subsequent release of proinflammatory cytokines and reactive oxygen species (ROS) [207,246,247]. In the presence of CNS injury, infection, or neurodegeneration, copious amounts of ATP are released into the extracellular environment from stressed and dying cells [248]. At high enough concentrations, this eATP can promote macrophage and microglial activation, which drives a cascade of P2X7R-mediated events that concludes with the time-dependent release of proinflammatory cytokines, including IL-6, IL-18, TNF-α, and IL-1β [59,247,249,250,251]. Specifically, P2X7R-mediated IL-1β release occurs through a multi-step process that requires priming first by cellular stress or pathogens (i.e., LPS) to stimulate the production of immature pro-IL-1β [121,122,252]. Upon accumulation in the cytosol, pro-IL-1β requires a secondary hit to promote its maturation and secretion driven by caspase-1 activity. Importantly, cells within a quiescent state store caspase-1 in an inactive form (pro-caspase-1), which requires cellular stimulation via DAMPs to undergo maturation to the active form [121]. In the case of eATP functioning as a DAMP, P2X7R activation drives significant K+ efflux from the intracellular environment, which subsequently promotes NLRP3 inflammasome complex formation [245,248,252]. The NLRP3 inflammasome is composed of a primary scaffold protein (NLRP3) that recruits the accessory protein ASC, which mediates pro-caspase-1 recruitment and activation [121,245]. Upon activation, caspase-1 drives the cleavage of pro-IL-1β into its active form, which can subsequently be secreted from the cell upon microvesicle shedding from the plasma membrane [13,241]. Importantly, P2X7R activation in microglia also promotes reactive oxygen species (ROS) formation [207,253]. Upon stimulation, P2X7R drives ROS production via p38 MAPK-dependent NADPH oxidase activation [207]. Notably, both the P2X7R-mediated release of proinflammatory cytokines and ROS are highly characterized in the pathophysiology of neurodegenerative diseases [248]. In Alzheimer’s Disease, the P2X7R is significantly upregulated adjacent to the characteristic Aβ plaques where surrounding activated microglial populations are colocalized [207]. Importantly, Aβ stimulates microglial activation, thereby driving the release of proinflammatory cytokines and ROS, which induce pro-apoptotic gene activity, thus mediating the death of neuronal cell populations and exacerbating neuroinflammation [248,254].

3.4. Tumor Microenvironment

Tumorigenesis frequently occurs at chronically inflamed tissue sites [255,256,257], where the rate at which the tumor proliferates is largely dependent on a delicate balance of immunosuppressive and immunostimulating cell types coexisting within the tumor microenvironment (TME) [63,191,258,259,260,261]. Underlying components of the TME include stromal cells, fibroblasts, endothelial cells, and infiltrating innate (TAMs, myeloid-derived suppressor cells, dendritic cells) and adaptive (T cells) immune cells [255,262]. Communication between cells occurs through the release of growth factors (VEGF), cytokines (IL-6), chemokines, components of the extracellular matrix, and purines to dictate tumor growth [255,257,263]. Tumor cells express P2X7Rs that drive proliferation by enhancing cellular metabolism and angiogenesis when the concentration of eATP is relatively low [260,264,265]. In contrast, when eATP rises to concentrations ≥100 µM as the result of hypoxic tissue necrosis, eATP promotes tumor cell cytotoxicity through sustained membrane permeabilization [63,260,266]. The ability of ATP to kill cancer cells justifies the development of selective P2X7R agonists as a therapeutic cancer target [266,267].
The immune cells that infiltrate tumors also express high densities of P2X7Rs, which in part determine whether these cells work to promote or eliminate tumor cells [256,268]. P2X7R-driven NLRP3 inflammasome complex activation and subsequent IL-1β release largely account for the immunostimulating qualities of P2X7R. Namely, IL-1β secretion from dendritic cells primes antigen-specific CD8+ T-cells, which release IFN-γ to exert their antitumor effects [256,269]. This tumor-eradicating role is exemplified in cancer models utilizing P2X7R-deficient mice (p2rx7−/−). In the absence of P2X7R, inoculated tumors both proliferated and metastasized at a faster rate compared to those brought up in wild-type mice (p2rx7+/+) [270]. Conversely, P2X7R activation on myeloid-derived suppressor cells fosters tumor-promotion upon the production and release of immunosuppressive factors, including reactive oxygen species, arginase-1, and TGF-β1 [256,263]. Additionally, P2X7R upregulation in glioma-associated microglia drives immunosuppression upon facilitating NLRP3 inflammasome activation and IL-1β release [48,268]. Proinflammatory IL-1β stimulates glioma cells to produce TGF-β, which mediates subsequent upregulation of VEGF, leading to tumor proliferation via increased angiogenesis [271,272,273]. Importantly, eATP also exhibits immunosuppressive effects upon its breakdown to adenosine via ectonucleotidases CD39 and CD73. Particularly, Tregs’ characteristic immunosuppressive activity is based on its high expression level of both ectonucleotidases. Free adenosine can then target P1 A2ARs to inhibit tumor-infiltrating cytotoxic CD8+ T cells [266,272].

3.5. Cell Death and Disease

Sustained stimulation with ATP is a potent catalytic stimulus for several cell types, including microglial cells, and the available literature clearly point towards the involvement of P2X7R in ATP-induced cell death, as reviewed by Peter Illes [274]. P2X7R has been described as a death receptor [43,275]. Short periods of P2X7R activation are cytotoxic, and once activated, the P2X7R sets in motion an irreversible death process [276,277]. Cells primed with inflammatory mediators (e.g., lipopolysaccharide) are particularly susceptible to the toxic actions of ATP [278], and this priming effect may alter the distribution or activation of P2X7 receptors in cell membranes [279].
Studies in culture strongly suggest a role for P2X7R-mediated cell death in a number of neurodegenerative diseases. Specific examples include rat microglial cell lines N9 and N13 [237], murine microglial cell line EOC13 [280], mouse primary microglia [207,277], and enteric glia [281].

3.6. Oxygen Glucose Deprivation

Oxygen glucose deprivation (OGD) is often used to study ischemic cell death. It negatively impacts microglia motility and induces microglia cell death. Upregulation of the P2X4R and P2X7R is reported to occur in N9 microglial cells deprived of oxygen [282]. Further, the same study suggests that metabolic stress like OGD induces massive release of extracellular ATP, which in turn activates cortical P2X and P2Y receptors. Several P2 receptors (P2X1R, P2X2R, P2X3R, P2X5R, and P2Y11R) alter the homeostatic balance of Ca2+ and Na+ fluxes, triggering both necrotic and apoptotic pathways [283]. In a similar fashion, P2X4 and P2X7 receptors induce the microglial release of proinflammatory cytokines [246] and subsequent neuronal death. Blocking the receptors with the P2 antagonists PPADS and TNP-ATP reduced microglia activation and rescued cortical cells from OGD-induced cell death [282]. OGD-induced microglial cell death has also been studied in BV2 cells, where the pharmacological inhibition of P2X7R using Brilliant Blue G (BBG) significantly reduced OGD-induced BV2 cell death. Similar results were observed in neonatal hippocampal slices. Here, OGD increases extracellular ATP, and treatments that decreased the concentration of extracellular ATP or reduced the availability of P2X7R receptors inhibited OGD-induced microglia cell death [284]. The depletion of extracellular Ca2+ also significantly inhibits cell death, indicating that OGD induces Ca2+-dependent microglia cell death [284]. Further in vivo studies performed on a middle cerebral artery occlusion rat model showed that inhibition of P2X7R expression by promoting degradation of ATP protects against the brain injury produced by OGD [285]. However, P2X7Rs are not the sole contributors to the purine- and calcium-dependent ischemic cell death and other mechanisms remain to be discovered.

4. Disease States

P2X7Rs are implicated in the genesis and pathology of a number of chronic diseases.

4.1. Alzheimer’s Disease

Alzheimer’s disease (AD) is a leading cause of dementia worldwide. No treatments are currently available to stop AD or significantly slow its progression. AD is characterized by cortical atrophy, neuroinflammation, the presence of extracellular senile plaques composed of amyloid beta peptides, intracellular neurofibrillary tangles, neuroinflammation, and the loss of neurons [286,287]. Microglia are implicated in AD, and phenotypic changes in microglia morphology, a hallmark of activation, are seen at the very start of disease progression, and microglia are widely acknowledged as a crucial target of Aβ-dependent toxicity. Further, the microglia of AD patients show an acute proinflammatory response and chronic reduction in phagocytic ability [288,289].
Several studies provide evidence that extracellular ATP plays an important role in ATP-induced neurotoxicity (Figure 2), primarily via the activation of P2X7R [290]. The upregulation of P2X7R has been observed in the microglia surrounding the senile plaques in both human and animal models of AD, such as the Tg2576 transgenic mice, APPswe/PS1De9, J20 mice, and rats following intrahippocampal Aβ injections [291].
The increased activation of P2X7R in microglia and ROS production concentrate in the area of senile plaques appears to increase with age and are parallel with Aβ increase and correlate with synaptotoxicity in AD [292]. In mouse microglia, Aβ increases intracellular Ca2+, promotes ATP release, permeabilizes the plasma membrane, and causes an P2X7R-dependent increase in IL-1β [293]. Furthermore, P2X7Rs also regulate Aβ dependent NLRP3 gene transcription, NLRP3 inflammasome activation, and release of IL-1β in microglia [251,294].
In both the in vivo and in vitro models, pharmacological blockade or genetic silencing of the P2X7R prevents activation of microglia and neuroinflammation induced by Aβ [295,296,297,298]. More importantly, in an AD mouse model induced by intrahippocampal Aβ injections, P2X7R also preserves memory and cognitive function by increasing the spine density of hippocampal neurons [299]. Similar effects were also demonstrated in APP/PS1 mice where a lack of P2X7R function led to an increase in synaptic plasticity and improved cognitive abilities [300]. In a familiar Alzheimer’s disease (FAD) model using transgenic P2X7REGFP/J20 mice and human AD patients, it has been shown that P2X7R is upregulated on microglia that surround the senile plaques [301]. Neuroinflammation caused by Aβ increased P2X7R expression on microglia in the advanced and late stages of AD, whereas its expression was found to be decreased in neurons. P2X7R activation enhances the migration of microglia towards senile plaque and reduces its phagocytic capability [301]. The silencing of P2X7R accelerates Aβ phagocytosis by microglia, reduces Aβ load, improves cognitive decline, rescues synaptic dysfunction, and decreases CD8+ T cell recruitment by reducing chemokine production [300].
The accumulation of Aβ also leads to the impairment of mitochondrial oxidative phosphorylation by altering complex IV, F0F1 ATP synthase and mitochondrial permeability transition pore (mPTP) [302,303,304]. P2X7R-deficient mice are resistant to Aβ dependent mitochondrial hyperpolarization and cyt c release [294]. The Aβ-dependent toxicity of microglia relies on the ability of P2X7R to promote cellular uptake of Aβ through innate phagocytosis and its localization to mitochondria. Thus, in isolated and permeabilized mitochondria, the lack of P2X7R was not able to prevent the inhibition of F0/F1 ATP synthase by Aβ. Further, this study showed that the Ca2+ channel blocker, nimodipine, strongly reduced all the Aβ induced effects, such as NF-κB activation, NLRP3 inflammasome stimulation, cyt c release IL1β release, mitochondrial dysfunction and cell death. Nimodipine did not exert its effect by directly antagonizing the P2X7R, but it prevented the delivery of Aβ to the mitochondria; thus, it might be possible that nimodipine targets a pathway downstream from the activation of P2X7R [294].
The 489C > T polymorphism in the P2X7R gene decreases the probability of having AD as this polymorphism alone or in concomitance with 1513A > C polymorphism is less frequent in AD patients compared to age-matched non-demented elderly [305]. A recent study performed on 5X FAD mice suggested that taurodeoxycholate (TDCA), a GPCR19 ligand, can reduce neuroinflammation in AD patients. In their study, TDCA was found to inhibit the priming phase of NLRP3 inflammasome activation, the suppression of P2X7R expression and P2X7R-mediated calcium mobilization, all of which are necessary for the production of IL-1β/IL-18 from microglia [306]. Transvagal nerve stimulation can be another promising strategy to improve spatial memory and learning in APP/PS1 mice by inhibiting hippocampal P2X7R/NLRP3/Caspase-1 signaling; however, more studies are required to establish its role in improving cognition [307]. Altogether, these studies suggest that modifying the microglia function or targeting the P2X7R may be a promising strategy in the therapeutic management of AD.

4.2. Parkinson’s Disease

Parkinson’s disease (PD) is a progressive nervous system disorder characterized by the loss of dopaminergic (DA) neurons in the substantia nigra and the presence of intra-neuronal cytoplasmic inclusions containing α-synuclein and neurites. Clinically, PD symptoms include bradykinesia, postural instability, resting tremors and muscle rigidity. In animal models of PD, such as 6-OHDA, MPTP and rotenone, and PD patient’s microglial activation and chronic neuroinflammation, have been shown to contribute to the degeneration of DA neurons in the striatum and substantia nigra [308,309]. The increased expression of P2X7R has been associated with PD and contributes to the pathogenesis of PD by affecting gliosis, synaptotoxicity and neurotoxicity [310,311,312]. The accumulation and oligomerization of misfolded α-synuclein named Lewy bodies has been recognized as a central player in the pathology of PD. In BV-2 cells and primary culture microglia, α-synuclein binds and activates P2X7R in microglia, activating the PI3K/AKT pathway and inducing ROS production [313]. This leads to deregulation in dopaminergic and glutamatergic synaptic transmission, causing neurotoxicity. α-synuclein also stimulates P2X7R transcription, which might be one of the reasons for the observed upregulation of P2X7R in PD patients. In mice, increased expression of P2X7R was observed with the acute 6-OHDA toxin model in a time-dependent manner, which coincides with an increase in translocator protein (TSPO), leading to increased binding of P2X7R with TSPO and motor behavior changes. Chronic α-synuclein model leads to an increase in TSPO without altering P2X7R expression, suggesting that increased P2X7R binding with TSPO is associated with neuroinflammation in acute but not chronic rodent models of Parkinson’s disease [314]. This study suggests that an alternative mechanism might play a role in inducing neuroinflammation relating to α-synucleinopathy in rodents. The microglia recognize, uptake and phagocyte α-synuclein. In rat LPS models and 6-OHDA models of PD, P2X7R antagonist BBG has been shown to reduce damage to dopaminergic neurons, attenuate LPS-induced upregulation of the expression of P2X7R, microglial activation, mitochondrial dysfunction and behavioral deficits [315,316]. Uncontrolled microglial activation in PD secretes Cathepsin L (CTSL) from microsomes, inducing neuronal damage and death. This study identified the P2X7R/PI3K/AKT signaling pathway as the underlying molecular mechanism [317]. The P2Y6 receptor has also been shown to have a potential role in the pathogenesis of PD. Apart from P2X7R, an increased level of P2Y6R has also been shown in the PBMCs of sporadic PD patients, suggesting that it is a potential clinical biomarker [318]. Both the in vivo and in vitro studies performed on the 6-OHDA model of PD have shown the neuroprotective effect of P2X7R and P2Y6 antagonists [319]. Thus, P2X7R and P2Y6R antagonists may have therapeutic potential in terms of PD.

4.3. Epilepsy

Epilepsy is a prevalent neurological disease that afflicts more than 70 million people worldwide [320], often presenting with cognitive, behavioral, and psychological abnormalities [321]. Current anti-epileptic drugs (AEDs) focus on a neurocentric approach to pharmacological intervention, thereby targeting the classical seizure model fixated on hypersynchronization of neuronal output and an imbalance in excitation/inhibition coupling [322,323]. However, thirty percent of epilepsies are completely refractive to AEDs [324], highlighting the importance of developing therapeutics with novel mechanisms of action. Importantly, recent studies suggest a strong contribution of neuroinflammation and neuroglia in the pathophysiology of epilepsy [325]. Specifically, translational work in mice and rats suggest that P2XRs could be novel targets for AED development due to epilepsy promoting the upregulation of P2Y6, P2Y12, P2X4, and P2X7 receptors on activated microglia [326,327] and promote neuroinflammation. Specifically, Wang et al. recently showed that usage of the anti-inflammatory drug Astaxanthin can attenuate the P2X7R-mediated neuroinflammation that is associated with status epilepticus (SE) [328]. As outlined above, limited release and rapid degradation limit the amount of ATP in the extracellular milieu of healthy brains. However, the concentration increases in response to the excessive neuronal firing that characteristically occurs throughout an epileptic seizure [326,327,329]. The result is a chronic neuroinflammatory state that disrupts the integrity of the blood–brain barrier, permitting infiltration of peripheral immune cells such as brain-infiltrating leukocytes into the brain, driving further inflammation [330,331]. A study performed using a lithium-pilocarpine-induced epileptic rat model demonstrates that increased expression of P2X7R in activated microglia promotes depression and anxiety-like behaviors in epileptic rats. P2X7R antagonist BBG reversed these effects as effectively as the classic anti-depressive and anti-anxiety drug fluoxetine [332]. Another study based on a similar animal model for status epilepticus (SE) revealed that protein disulfide isomerase (PDI)-mediated redox/S-nitrosylation may be responsible for facilitating the trafficking of P2X7R to the cell surface. P2X7R further promotes microglial activation and astroglial apoptosis following SE as L-NAME and PDI siRNA-attenuated SE-induced microglial activation and astroglial apoptosis [333]. In palmitoyl protein thioesterase 1 (PPT1) knock-in mice, astrocyte activation precedes microglial activation and neuronal death. The decreased expression of synaptic protein GluN2B and GABAARα1 in the hippocampus of PPT1-KI mice was observed, along with the activation of P2X7R in the microglia, causing the release of proinflammatory cytokines IL-1β and TNF-α. The cytokines further act on neurons and astrocytes, releasing ATP and glutamate, leading to increased neuronal excitability and seizures [334]. Such a scenario suggests that targeting P2X7R-mediated microglial activation and the subsequent release of proinflammatory cytokines may limit epilepsy-associated neuroinflammation, providing a novel mechanism of action to treat drug-resistant epilepsy. Treatment with P2X7 receptor antagonists has also proven to be beneficial in hypoxia-induced neonatal seizures and the subsequent development of epilepsy [335].

5. Conclusions

While microglial P2X7Rs undoubtedly play key roles in the formation, maintenance, and pathology of the CNS, our understanding of how this neuromodulation occurs is incompletely understood. This is particularly true of the actions of eATP on human microglia, for which an accessible and reliable in vitro experimental model is lacking. Future studies of human microglia maintained in a milieu that accurately replicates the CNS microenvironments are badly needed. Until then, we are left to rely on increasingly sophisticated animal models that regularly highlight the importance of microglia, eATP, and P2X7Rs in initiating and modulating CNS function.

Author Contributions

M.T., S.M. and T.M.E. equally contributed to the conceptualization, drafting, and reviewing of this manuscript. All authors agree to be accountable for the accuracy and integrity of the work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by a grant from the National Institutes of Health (R01GM112188 to T.M.E.).

Acknowledgments

We thank Peter Illes for constructive comments on an earlier version of this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rajendran, M.; Dane, E.; Conley, J.; Tantama, M. Imaging Adenosine Triphosphate (ATP). Biol. Bull. 2016, 231, 73–84. [Google Scholar] [CrossRef]
  2. Khakh, B.S.; Burnstock, G. The Double Life of ATP. Sci. Am. 2009, 301, 84–92. [Google Scholar] [CrossRef] [PubMed]
  3. Giuliani, A.L.; Sarti, A.C.; Di Virgilio, F. Extracellular Nucleotides and Nucleosides as Signalling Molecules. Immunol. Lett. 2019, 205, 16–24. [Google Scholar] [CrossRef] [PubMed]
  4. Burnstock, G. A Basis for Distinguishing Two Types of Purinergic Receptor. In Cell Membrane Receptors for Drugs and Hormones A Multidisciplinary Approach; Raven Press: New York, NY, USA, 1978. [Google Scholar]
  5. Burnstock, G.; Kennedy, C. Is There a Basis for Distinguishing Two Types of P2-Purinoceptor? Gen. Pharmacol. 1985, 16, 433–440. [Google Scholar] [CrossRef] [PubMed]
  6. Hammarberg, C.; Schulte, G.; Fredholm, B.B. Evidence for Functional Adenosine A3 Receptors in Microglia Cells. J. Neurochem. 2003, 86, 1051–1054. [Google Scholar] [CrossRef]
  7. Synowitz, M.; Glass, R.; Färber, K.; Markovic, D.; Kronenberg, G.; Herrmann, K.; Schnermann, J.; Nolte, C.; van Rooijen, N.; Kiwit, J.; et al. A1 Adenosine Receptors in Microglia Control Glioblastoma-Host Interaction. Cancer Res. 2006, 66, 8550–8557. [Google Scholar] [CrossRef] [PubMed]
  8. Orr, A.G.; Orr, A.L.; Li, X.-J.; Gross, R.E.; Traynelis, S.F.; Neurosci, N. Adenosine A 2A Receptor Mediates Microglial Process Retraction HHS Public Access Author Manuscript. Nat. Neurosci. 2009, 12, 872–878. [Google Scholar] [CrossRef]
  9. Haselkorn, M.L.; Shellington, D.K.; Jackson, E.K.; Vagni, V.A.; Janesko-Feldman, K.; Dubey, R.K.; Gillespie, D.G.; Cheng, D.; Bell, M.J.; Jenkins, L.W.; et al. Adenosine A1 Receptor Activation as a Brake on the Microglial Response after Experimental Traumatic Brain Injury in Mice. J. Neurotrauma 2010, 27, 901–910. [Google Scholar] [CrossRef]
  10. Koscsó, B.; Csóka, B.; Selmeczy, Z.; Himer, L.; Pacher, P.; Virág, L.; Haskó, G. Adenosine Augments IL-10 Production by Microglial Cells through an A2B Adenosine Receptor-Mediated Process. J. Immunol. 2012, 188, 445–453. [Google Scholar] [CrossRef]
  11. Ohsawa, K.; Sanagi, T.; Nakamura, Y.; Suzuki, E.; Inoue, K.; Kohsaka, S. Adenosine A3 Receptor is Involved in ADP-Induced Microglial Process Extension and Migration. J. Neurochem. 2012, 121, 217–227. [Google Scholar] [CrossRef]
  12. Merighi, S.; Bencivenni, S.; Vincenzi, F.; Varani, K.; Borea, P.A.; Gessi, S. A 2B Adenosine Receptors Stimulate IL-6 Production in Primary Murine Microglia through p38 MAPK Kinase Pathway. Pharmacol. Res. 2017, 117, 9–19. [Google Scholar] [CrossRef] [PubMed]
  13. Bianco, F.; Pravettoni, E.; Colombo, A.; Schenk, U.; Möller, T.; Matteoli, M.; Verderio, C. Astrocyte-Derived ATP Induces Vesicle Shedding and IL-1β Release from Microglia. J. Immunol. 2005, 174, 7268–7277. [Google Scholar] [CrossRef] [PubMed]
  14. Xiang, Z.; Burnstock, G. Expression of P2X Receptors on Rat Microglial Cells during Early Development. Glia 2005, 52, 119–126. [Google Scholar] [CrossRef] [PubMed]
  15. Calovi, S.; Mut-Arbona, P.; Sperlágh, B. Microglia and the Purinergic Signaling System. Neuroscience 2019, 405, 137–147. [Google Scholar] [CrossRef] [PubMed]
  16. von Kügelgen, I. Pharmacology of P2Y Receptors. Brain Res. Bull. 2019, 151, 12–24. [Google Scholar] [CrossRef]
  17. Hidetoshi, T.-S.; Makoto, T.; Inoue, K. P2Y Receptors in Microglia and Neuroinflammation. Wiley Interdiscip. Rev. Membr. Transp. Signal. 2012, 1, 493–501. [Google Scholar] [CrossRef]
  18. Swiatkowski, P.; Murugan, M.; Eyo, U.; Wang, Y.; Rangaraju, S.; Oh, S.; Wu, L.-J. Activation of Microglial P2Y12 Receptor Is Required for Outward Potassium Currents in Response to Neuronal Injury. Neuroscience 2016, 318, 22–33. [Google Scholar] [CrossRef]
  19. Madry, C.; Kyrargyri, V.; Arancibia-Cárcamo, I.L.; Jolivet, R.; Kohsaka, S.; Bryan, R.M.; Attwell, D. Microglial Ramification, Surveillance, and Interleukin-1β Release Are Regulated by the Two-Pore Domain K+ Channel THIK-1. Neuron 2018, 97, 299–312.e6. [Google Scholar] [CrossRef]
  20. Illes, P.; Rubini, P.; Ulrich, H.; Zhao, Y.; Tang, Y. Regulation of Microglial Functions by Purinergic Mechanisms in the Healthy and Diseased CNS. Cells 2020, 9, 1108. [Google Scholar] [CrossRef]
  21. Egan, T.M.; Samways, D.S.K.; Li, Z. Biophysics of P2X Receptors. Pflüg. Arch. Eur. J. Physiol. 2006, 452, 501–512. [Google Scholar] [CrossRef]
  22. Lê, K.-T.; Paquet, M.; Nouel, D.; Babinski, K.; Séguéla, P. Primary Structure and Expression of a Naturally Truncated Human P2X ATP Receptor Subunit from Brain and Immune System. FEBS Lett. 1997, 418, 195–199. [Google Scholar] [CrossRef] [PubMed]
  23. Bo, X.; Jiang, L.-H.; Wilson, H.L.; Kim, M.; Burnstock, G.; Surprenant, A.; North, R.A. Pharmacological and Biophysical Properties of the Human P2X5 Receptor. Mol. Pharmacol. 2003, 63, 1407–1416. [Google Scholar] [CrossRef] [PubMed]
  24. Ormond, S.J.; Barrera, N.P.; Qureshi, O.S.; Henderson, R.M.; Edwardson, J.M.; Murrell-Lagnado, R.D. An Uncharged Region within the N Terminus of the P2X6 Receptor Inhibits Its Assembly and Exit from the Endoplasmic Reticulum. Mol. Pharmacol. 2006, 69, 1692–1700. [Google Scholar] [CrossRef]
  25. Murrell-Lagnado, R.D.; Qureshi, O.S. Assembly and Trafficking of P2X Purinergic Receptors (Review). Mol. Membr. Biol. 2008, 25, 321–331. [Google Scholar] [CrossRef] [PubMed]
  26. Torres, G.E.; Egan, T.M.; Voigt, M.M. Hetero-Oligomeric Assembly of P2X Receptor Subunits. Specificities Exist with Regard to Possible Partners. J. Biol. Chem. 1999, 274, 6653–6659. [Google Scholar] [CrossRef] [PubMed]
  27. Saul, A.; Hausmann, R.; Kless, A.; Nicke, A. Heteromeric Assembly of P2X Subunits. Front. Cell. Neurosci. 2013, 7, 250. [Google Scholar] [CrossRef] [PubMed]
  28. Nicke, A. Homotrimeric Complexes Are the Dominant Assembly State of Native P2X7 Subunits. Biochem. Biophys. Res. Commun. 2008, 377, 803–808. [Google Scholar] [CrossRef]
  29. Cheewatrakoolpong, B.; Gilchrest, H.; Anthes, J.C.; Greenfeder, S. Identification and Characterization of Splice Variants of the Human P2X7 ATP Channel. Biochem. Biophys. Res. Commun. 2005, 332, 17–27. [Google Scholar] [CrossRef]
  30. Sluyter, R. The P2X7 Receptor. Adv. Exp. Med. Biol. 2017, 19, 17–53. [Google Scholar]
  31. Liang, X.; Samways, D.S.K.; Wolf, K.; Bowles, E.A.; Richards, J.P.; Bruno, J.; Dutertre, S.; DiPaolo, R.J.; Egan, T.M. Quantifying Ca2+ Current and Permeability in ATP-Gated P2X7 Receptors. J. Biol. Chem. 2015, 290, 7930–7942. [Google Scholar] [CrossRef]
  32. Surprenant, A.; Rassendren, F.; Kawashima, E.; North, R.A.; Buell, G. The Cytolytic P2Z Receptor for Extracellular ATP Identified as a P2X Receptor (P2X7). Science 1996, 272, 735–738. [Google Scholar] [CrossRef] [PubMed]
  33. Habermacher, C.; Dunning, K.; Chataigneau, T.; Grutter, T. Molecular Structure and Function of P2X Receptors. Neuropharmacology 2016, 104, 18–30. [Google Scholar] [CrossRef] [PubMed]
  34. McCarthy, A.E.; Yoshioka, C.; Mansoor, S.E. Full-Length P2X7 Structures Reveal How Palmitoylation Prevents Channel Desensitization. Cell 2019, 179, 659–670.e13. [Google Scholar] [CrossRef] [PubMed]
  35. Costa-Junior, H.M.; Vieira, F.S.; Coutinho-Silva, R. C Terminus of the P2X7 Receptor: Treasure Hunting. Purinergic Signal. 2011, 7, 7–19. [Google Scholar] [CrossRef] [PubMed]
  36. Kanellopoulos, J.M.; Delarasse, C. Pleiotropic Roles of P2X7 in the Central Nervous System. Front. Cell. Neurosci. 2019, 13, 401. [Google Scholar] [CrossRef] [PubMed]
  37. Kopp, R.; Krautloher, A.; Ramírez-Fernández, A.; Nicke, A. P2X7 Interactions and Signaling—Making Head or Tail of It. Front. Mol. Neurosci. 2019, 12, 183. [Google Scholar] [CrossRef] [PubMed]
  38. Jiang, L.-H. Inhibition of P2X7 Receptors by Divalent Cations: Old Action and New Insight. Eur. Biophys. J. 2009, 38, 339–346. [Google Scholar] [CrossRef] [PubMed]
  39. Nicke, A.; Kuan, Y.-H.; Masin, M.; Rettinger, J.; Marquez-Klaka, B.; Bender, O.; Górecki, D.C.; Murrell-Lagnado, R.D.; Soto, F. A Functional P2X7 Splice Variant with an Alternative Transmembrane Domain 1 Escapes Gene Inactivation in P2X7 Knock-out Mice. J. Biol. Chem. 2009, 284, 25813–25822. [Google Scholar] [CrossRef]
  40. Zimmermann, H. Extracellular Metabolism of ATP and Other Nucleotides. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2000, 362, 299–309. [Google Scholar] [CrossRef]
  41. Di Virgilio, F.; Boeynaems, J.-M.; Robson, S.C. Extracellular Nucleotides as Negative Modulators of Immunity. Curr. Opin. Pharmacol. 2009, 9, 507–513. [Google Scholar] [CrossRef]
  42. Trautmann, A. Extracellular ATP in the Immune System: More than Just a “Danger Signal”. Sci. Signal. 2009, 2, pe6. [Google Scholar] [CrossRef] [PubMed]
  43. Di Virgilio, F. ATP as a Death Factor. BioFactors 1998, 8, 301–303. [Google Scholar] [CrossRef]
  44. Linden, J.; Koch-Nolte, F.; Dahl, G. Purine Release, Metabolism, and Signaling in the Inflammatory Response. Annu. Rev. Immunol. 2019, 37, 325–347. [Google Scholar] [CrossRef] [PubMed]
  45. Gordon, J.L. Extracellular ATP: Effects, Sources and Fate. Biochem. J. 1986, 233, 309–319. [Google Scholar] [CrossRef]
  46. Guerra, A.N.; Gavala, M.L.; Chung, H.S.; Bertics, P.J. Nucleotide Receptor Signalling and the Generation of Reactive Oxygen Species. Purinergic Signal. 2007, 3, 39–51. [Google Scholar] [CrossRef] [PubMed]
  47. Pellegatti, P.; Raffaghello, L.; Bianchi, G.; Piccardi, F.; Pistoia, V.; Di Virgilio, F. Increased Level of Extracellular ATP at Tumor Sites: In Vivo Imaging with Plasma Membrane Luciferase. PLoS ONE 2008, 3, e2599. [Google Scholar] [CrossRef]
  48. McLarnon, J.G. Roles of Purinergic P2X7 Receptor in Glioma and Microglia in Brain Tumors. Cancer Lett. 2017, 402, 93–99. [Google Scholar] [CrossRef]
  49. Morciano, G.; Sarti, A.C.; Marchi, S.; Missiroli, S.; Falzoni, S.; Raffaghello, L.; Pistoia, V.; Giorgi, C.; Di Virgilio, F.; Pinton, P. Use of Luciferase Probes to Measure ATP in Living Cells and Animals. Nat. Protoc. 2017, 12, 1542–1562. [Google Scholar] [CrossRef]
  50. Rodrigues, R.J.; Tomé, A.R.; Cunha, R.A. ATP as a Multi-Target Danger Signal in the Brain. Front. Neurosci. 2015, 9, 148. [Google Scholar] [CrossRef]
  51. Honda, S.; Sasaki, Y.; Ohsawa, K.; Imai, Y.; Nakamura, Y.; Inoue, K.; Kohsaka, S. Extracellular ATP or ADP Induce Chemotaxis of Cultured Microglia through Gi/o-Coupled P2Y Receptors. J. Neurosci. 2001, 21, 1975–1982. [Google Scholar] [CrossRef]
  52. Haynes, S.E.; Hollopeter, G.; Yang, G.; Kurpius, D.; Dailey, M.E.; Gan, W.-B.; Julius, D. The P2Y12 Receptor Regulates Microglial Activation by Extracellular Nucleotides. Nat. Neurosci. 2006, 9, 1512–1519. [Google Scholar] [CrossRef] [PubMed]
  53. Elliott, M.R.; Chekeni, F.B.; Trampont, P.C.; Lazarowski, E.R.; Kadl, A.; Walk, S.F.; Park, D.; Woodson, R.I.; Ostankovich, M.; Sharma, P.; et al. Nucleotides Released by Apoptotic Cells Act as a Find-Me Signal to Promote Phagocytic Clearance. Nature 2009, 461, 282–286. [Google Scholar] [CrossRef]
  54. Ravichandran, K.S. Beginnings of a Good Apoptotic Meal: The Find-Me and Eat-Me Signaling Pathways. Immunity 2011, 35, 445–455. [Google Scholar] [CrossRef] [PubMed]
  55. von Kügelgen, I. Structure, Pharmacology and Roles in Physiology of the P2Y12 Receptor. In Advances in Experimental Medicine and Biology; Springer: Berlin/Heidelberg, Germany, 2017; Volume 1051. [Google Scholar] [CrossRef]
  56. Courageot, M.-P.; Lépine, S.; Hours, M.; Giraud, F.; Sulpice, J.-C. Involvement of Sodium in Early Phosphatidylserine Exposure and Phospholipid Scrambling Induced by P2X7 Purinoceptor Activation in Thymocytes. J. Biol. Chem. 2004, 279, 21815–21823. [Google Scholar] [CrossRef] [PubMed]
  57. Sluyter, R.; Shemon, A.N.; Wiley, J.S. P2X7 Receptor Activation Causes Phosphatidylserine Exposure in Human Erythrocytes. Biochem. Biophys. Res. Commun. 2007, 355, 169–173. [Google Scholar] [CrossRef] [PubMed]
  58. Zhang, X.; Meng, L.; He, B.; Chen, J.; Liu, P.; Zhao, J.; Zhang, Y.; Li, M.; An, D. The Role of P2X7 Receptor in ATP-Mediated Human Leukemia Cell Death: Calcium Influx-Independent. Acta Biochim. Biophys. Sin. 2009, 41, 362–369. [Google Scholar] [CrossRef]
  59. Janks, L.; Sharma, C.V.R.; Egan, T.M. A Central Role for P2X7 Receptors in Human Microglia. J. Neuroinflamm. 2018, 15, 325. [Google Scholar] [CrossRef]
  60. Gu, B.J.; Wiley, J.S. P2X7 as a Scavenger Receptor for Innate Phagocytosis in the Brain. Br. J. Pharmacol. 2018, 175, 4195–4208. [Google Scholar] [CrossRef]
  61. Gu, B.J.; Zhang, W.Y.; Bendall, L.J.; Chessell, I.P.; Buell, G.N.; Wiley, J.S. Expression of P2X7 Purinoceptors on Human Lymphocytes and Monocytes: Evidence for Nonfunctional P2X7 receptors. Am. J. Physiol. Cell Physiol. 2000, 279, C1189–C1197. [Google Scholar] [CrossRef]
  62. Barden, J.A. Non-Functional P2X7: A Novel and Ubiquitous Target in Human Cancer. J. Clin. Cell. Immunol. 2014, 5, 4. [Google Scholar] [CrossRef]
  63. Gilbert, S.; Oliphant, C.; Hassan, S.; Peille, A.; Bronsert, P.; Falzoni, S.; Di Virgilio, F.; McNulty, S.; Lara, R. ATP in the Tumour Microenvironment Drives Expression of nfP2X7, a Key Mediator of Cancer Cell Survival. Oncogene 2019, 38, 194–208. [Google Scholar] [CrossRef]
  64. Azevedo, F.A.C.; Carvalho, L.R.B.; Grinberg, L.T.; Farfel, J.M.; Ferretti, R.E.L.; Leite, R.E.P.; Filho, W.J.; Lent, R.; Herculano-Houzel, S. Equal Numbers of Neuronal and Nonneuronal Cells Make the Human Brain an Isometrically Scaled-Up Primate Brain. J. Comp. Neurol. 2009, 513, 532–541. [Google Scholar] [CrossRef]
  65. Fan, X.; Agid, Y. At the Origin of the History of Glia. Neuroscience 2018, 385, 255–271. [Google Scholar] [CrossRef]
  66. Allen, N.J.; Lyons, D.A. Glia as Architects of Central Nervous System Formation and Function. Science 2018, 362, 181–185. [Google Scholar] [CrossRef]
  67. Li, Q.; Barres, B.A. Microglia and Macrophages in Brain Homeostasis and Disease. Nat. Rev. Immunol. 2018, 18, 225–242. [Google Scholar] [CrossRef]
  68. 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]
  69. Ginhoux, F.; Greter, M.; Leboeuf, M.; Nandi, S.; See, P.; Gokhan, S.; Mehler, M.F.; Conway, S.J.; Ng, L.G.; Stanley, E.R.; et al. Fate Mapping Analysis Reveals That Adult Microglia Derive from Primitive Macrophages. Science 2010, 330, 841–845. [Google Scholar] [CrossRef]
  70. Brioschi, S.; Zhou, Y.; Colonna, M. Brain Parenchymal and Extraparenchymal Macrophages in Development, Homeostasis, and Disease. J. Immunol. 2020, 204, 294–305. [Google Scholar] [CrossRef]
  71. Verney, C.; Monier, A.; Fallet-Bianco, C.; Gressens, P. Early Microglial Colonization of the Human Forebrain and Possible Involvement in Periventricular White-Matter Injury of Preterm Infants. J. Anat. 2010, 217, 436–448. [Google Scholar] [CrossRef]
  72. Menassa, D.A.; Gomez-Nicola, D. Microglial Dynamics during Human Brain Development. Front. Immunol. 2018, 9, 1014. [Google Scholar] [CrossRef]
  73. Smolders, S.M.-T.; Kessels, S.; Vangansewinkel, T.; Rigo, J.-M.; Legendre, P.; Brône, B. Microglia: Brain Cells on the Move. Prog. Neurobiol. 2019, 178, 101612. [Google Scholar] [CrossRef] [PubMed]
  74. Bennett, F.C.; Bennett, M.L.; Yaqoob, F.; Mulinyawe, S.B.; Grant, G.A.; Gephart, M.H.; Plowey, E.D.; Barres, B.A. A Combination of Ontogeny and CNS Environment Establishes Microglial Identity. Neuron 2018, 98, 1170–1183.e8. [Google Scholar] [CrossRef]
  75. Li, Q.; Cheng, Z.; Zhou, L.; Darmanis, S.; Neff, N.F.; Okamoto, J.; Gulati, G.; Bennett, M.L.; Sun, L.O.; Clarke, L.E.; et al. Developmental Heterogeneity of Microglia and Brain Myeloid Cells Revealed by Deep Single-Cell RNA Sequencing. Neuron 2019, 101, 207–223.e10. [Google Scholar] [CrossRef]
  76. Hickman, S.E.; 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]
  77. Lai, A.Y.; Dhami, K.S.; Dibal, C.D.; Todd, K.G. Neonatal Rat Microglia Derived from different Brain Regions Have Distinct Activation Responses. Neuron Glia Biol. 2012, 7, 5–16. [Google Scholar] [CrossRef]
  78. Grabert, K.; Michoel, T.; Karavolos, M.H.; Clohisey, S.; Baillie, J.K.; Stevens, M.P.; Freeman, T.C.; Summers, K.M.; McColl, B.W. Microglial Brain Region–Dependent Diversity and Selective Regional Sensitivities to Aging. Nat. Neurosci. 2016, 19, 504–516. [Google Scholar] [CrossRef]
  79. De Biase, L.M.; Schuebel, K.E.; Fusfeld, Z.H.; Jair, K.; Hawes, I.A.; Cimbro, R.; Zhang, H.-Y.; Liu, Q.-R.; Shen, H.; Xi, Z.-X.; et al. Local Cues Establish and Maintain Region-Specific Phenotypes of Basal Ganglia Microglia. Neuron 2017, 95, 341–356.e6. [Google Scholar] [CrossRef]
  80. Ayata, P.; Badimon, A.; Strasburger, H.J.; Duff, M.K.; Montgomery, S.E.; Loh, Y.-H.E.; Ebert, A.; Pimenova, A.A.; Ramirez, B.R.; Chan, A.T.; et al. Epigenetic Regulation of Brain Region-Specific Microglia Clearance Activity. Nat. Neurosci. 2018, 21, 1049–1060. [Google Scholar] [CrossRef]
  81. Tan, Y.-L.; Yuan, Y.; Tian, L. Microglial Regional Heterogeneity and Its Role in the Brain. Mol. Psychiatry 2020, 25, 351–367. [Google Scholar] [CrossRef]
  82. Böttcher, C.; Psy, N.; Schlickeiser, S.; Sneeboer, M.A.M.; Kunkel, D.; Knop, A.; Paza, E.; Fidzinski, P.; Kraus, L.; Snijders, G.J.L.; et al. Human Microglia Regional Heterogeneity and Phenotypes Determined by Multiplexed Single-Cell Mass Cytometry. Nat. Neurosci. 2019, 22, 78–90. [Google Scholar] [CrossRef]
  83. Masuda, T.; Sankowski, R.; Staszewski, O.; Böttcher, C.; Amann, L.; Sagar, S.; Scheiwe, C.; Nessler, S.; Kunz, P.; Van Loo, G.; et al. Spatial and Temporal Heterogeneity of Mouse and Human Microglia at Single-Cell Resolution. Nature 2019, 566, 388–392. [Google Scholar] [CrossRef] [PubMed]
  84. van der Poel, M.; Ulas, T.; Mizee, M.R.; Hsiao, C.-C.; Miedema, S.S.M.; Adelia, N.; Schuurman, K.G.; Helder, B.; Tas, S.W.; Schultze, J.L.; et al. Transcriptional Profiling of Human Microglia Reveals Grey–White Matter Heterogeneity and Multiple Sclerosis-Associated Changes. Nat. Commun. 2019, 10, 1139. [Google Scholar] [CrossRef]
  85. Schwarz, J.M.; Sholar, P.W.; Bilbo, S.D. Sex Differences in Microglial Colonization of the Developing Rat Brain. J. Neurochem. 2012, 120, 948–963. [Google Scholar] [CrossRef]
  86. Guneykaya, D.; Ivanov, A.; Hernandez, D.P.; Haage, V.; Wojtas, B.; Meyer, N.; Maricos, M.; Jordan, P.; Buonfiglioli, A.; Gielniewski, B.; et al. Transcriptional and Translational Differences of Microglia from Male and Female Brains. Cell Rep. 2018, 24, 2773–2783.e6. [Google Scholar] [CrossRef] [PubMed]
  87. Kodama, L.; Gan, L. Do Microglial Sex Differences Contribute to Sex Differences in Neurodegenerative Diseases? Trends Mol. Med. 2019, 25, 741–749. [Google Scholar] [CrossRef] [PubMed]
  88. Villa, A.; Della Torre, S.; Maggi, A. Sexual Differentiation of Microglia. Front. Neuroendocr. 2019, 52, 156–164. [Google Scholar] [CrossRef] [PubMed]
  89. Hanamsagar, R.; Alter, M.D.; Block, C.S.; Sullivan, H.; Bolton, J.L.; Bilbo, S.D. Generation of a Microglial Developmental Index in Mice and in Humans Reveals a Sex Difference in Maturation and Immune Reactivity. Glia 2017, 65, 1504–1520. [Google Scholar] [CrossRef]
  90. Ginhoux, F.; Lim, S.; Hoeffel, G.; Low, D.; Huber, T. Origin and Differentiation of Microglia. Front. Cell Neurosci. 2013, 7, 45. [Google Scholar] [CrossRef]
  91. Askew, K.; Li, K.; Olmos-Alonso, A.; Garcia-Moreno, F.; Liang, Y.; Richardson, P.; Tipton, T.; Chapman, M.A.; Riecken, K.; Beccari, S.; et al. Coupled Proliferation and Apoptosis Maintain the Rapid Turnover of Microglia in the Adult Brain. Cell Rep. 2017, 18, 391–405. [Google Scholar] [CrossRef]
  92. Huang, Y.; Xu, Z.; Xiong, S.; Sun, F.; Qin, G.; Hu, G.; Wang, J.; Zhao, L.; Liang, Y.-X.; Wu, T.; et al. Repopulated Microglia Are Solely Derived from the Proliferation of Residual Microglia after Acute Depletion. Nat. Neurosci. 2018, 21, 530–540. [Google Scholar] [CrossRef]
  93. Dubbelaar, M.L.; Kracht, L.; Eggen, B.J.L.; Boddeke, E.W.G.M. The Kaleidoscope of Microglial Phenotypes. Front. Immunol. 2018, 9, 1753. [Google Scholar] [CrossRef]
  94. Tay, T.L.; Mai, D.; Dautzenberg, J.; Fernandez-Klett, F.; Lin, G.; Sagar, S.; Datta, M.; Drougard, A.; Stempfl, T.; Ardura-Fabregat, A.; et al. A New Fate Mapping System Reveals Context-Dependent Random or Clonal Expansion of Microglia. Nat. Neurosci. 2017, 20, 793–803. [Google Scholar] [CrossRef] [PubMed]
  95. Réu, P.; Khosravi, A.; Bernard, S.; Mold, J.E.; Salehpour, M.; Alkass, K.; Perl, S.; Tisdale, J.; Possnert, G.; Druid, H.; et al. The Lifespan and Turnover of Microglia in the Human Brain. Cell Rep. 2017, 20, 779–784. [Google Scholar] [CrossRef] [PubMed]
  96. Prinz, M.; Priller, J.; Sisodia, S.S.; Ransohoff, R.M. Heterogeneity of CNS Myeloid Cells and Their Roles in Neurodegeneration. Nat. Neurosci. 2011, 14, 1227–1235. [Google Scholar] [CrossRef]
  97. Bennett, M.L.; Bennett, F.C.; Liddelow, S.A.; Ajami, B.; Zamanian, J.L.; Fernhoff, N.B.; Mulinyawe, S.B.; Bohlen, C.J.; Adil, A.; Tucker, A.; et al. New Tools for Studying Microglia in the Mouse and Human CNS. Proc. Natl. Acad. Sci. USA 2016, 113, E1738–E1746. [Google Scholar] [CrossRef] [PubMed]
  98. Antel, J.P.; Becher, B.; Ludwin, S.K.; Prat, A.; Quintana, F.J. Glial Cells as Regulators of Neuroimmune Interactions in the Central Nervous System. J. Immunol. 2020, 204, 251–255. [Google Scholar] [CrossRef]
  99. Senatorov, V.V.; Friedman, A.R.; Milikovsky, D.Z.; Ofer, J.; Saar-Ashkenazy, R.; Charbash, A.; Jahan, N.; Chin, G.; Mihaly, E.; Lin, J.M.; et al. Blood-Brain Barrier Dysfunction in Aging Induces Hyperactivation of TGFβ Signaling and Chronic Yet Reversible Neural Dysfunction. Sci. Transl. Med. 2019, 11, eaaw8283. [Google Scholar] [CrossRef]
  100. Bohlen, C.J.; Bennett, F.C.; Tucker, A.F.; Collins, H.Y.; Mulinyawe, S.B.; Barres, B.A. Diverse Requirements for Microglial Survival, Specification, and Function Revealed by Defined-Medium Cultures. Neuron 2017, 94, 759–773.e8. [Google Scholar] [CrossRef]
  101. Geirsdottir, L.; David, E.; Keren-Shaul, H.; Weiner, A.; Bohlen, S.C.; Neuber, J.; Balic, A.; Giladi, A.; Sheban, F.; Dutertre, C.-A.; et al. Cross-Species Single-Cell Analysis Reveals Divergence of the Primate Microglia Program. Cell 2019, 179, 1609–1622.e16. [Google Scholar] [CrossRef]
  102. Smith, A.M.; Dragunow, M. The Human Side of Microglia. Trends Neurosci. 2014, 37, 125–135. [Google Scholar] [CrossRef]
  103. Volonté, C.; D’ambrosi, N. Membrane compartments and purinergic signalling: The Purinome, a Complex Interplay among Ligands, Degrading Enzymes, Receptors and Transporters. FEBS J. 2009, 276, 318–329. [Google Scholar] [CrossRef]
  104. Burnstock, G. Introduction to Purinergic Signalling in the Brain. In Advances in Experimental Medicine and Biology; Springer: Berlin/Heidelberg, Germany, 2020; Volume 1202. [Google Scholar] [CrossRef]
  105. Zimmermann, H.; Zebisch, M.; Sträter, N. Cellular Function and Molecular Structure of Ecto-Nucleotidases. Purinergic Signal. 2012, 8, 437–502. [Google Scholar] [CrossRef] [PubMed]
  106. Yegutkin, G.G. Enzymes Involved in Metabolism of Extracellular Nucleotides and Nucleosides: Functional Implications and Measurement of Activities. Crit. Rev. Biochem. Mol. Biol. 2014, 49, 473–497. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, Y.; Chen, K.; Sloan, S.A.; Bennett, M.L.; Scholze, A.R.; O’Keeffe, S.; Phatnani, H.P.; Guarnieri, P.; Caneda, C.; Ruderisch, N.; et al. An RNA-Sequencing Transcriptome and Splicing Database of Glia, Neurons, and Vascular Cells of the Cerebral Cortex. J. Neurosci. 2014, 34, 11929–11947. [Google Scholar] [CrossRef] [PubMed]
  108. He, Y.; Taylor, N.; Fourgeaud, L.; Bhattacharya, A. The Role of Microglial P2X7: Modulation of Cell Death and Cytokine Release. J. Neuroinflamm. 2017, 14, 135. [Google Scholar] [CrossRef]
  109. Sperlágh, B.; Illes, P. Purinergic Modulation of Microglial Cell Activation. Purinergic Signal. 2007, 3, 117–127. [Google Scholar] [CrossRef]
  110. Domercq, M.; Vázquez-Villoldo, N.; Matute, C. Neurotransmitter Signaling in the Pathophysiology of Microglia. Front. Cell. Neurosci. 2013, 7, 49. [Google Scholar] [CrossRef]
  111. Luongo, L.; Guida, F.; Imperatore, R.; Napolitano, F.; Gatta, L.; Cristino, L.; Giordano, C.; Siniscalco, D.; Di Marzo, V.; Bellini, G.; et al. The A1 Adenosine Receptor as a New Player in Microglia Physiology. Glia 2014, 62, 122–132. [Google Scholar] [CrossRef]
  112. Milior, G.; Morin-Brureau, M.; Chali, F.; Le Duigou, C.; Savary, E.; Huberfeld, G.; Rouach, N.; Pallud, J.; Capelle, L.; Navarro, V.; et al. Distinct P2Y Receptors Mediate Extension and Retraction of Microglial Processes in Epileptic and Peritumoral Human Tissue. J. Neurosci. 2020, 40, 1373–1388. [Google Scholar] [CrossRef]
  113. Avignone, E.; Ulmann, L.; Levavasseur, F.; Rassendren, F.; Audinat, E. Status Epilepticus Induces a Particular Microglial Activation State Characterized by Enhanced Purinergic Signaling. J. Neurosci. 2008, 28, 9133–9144. [Google Scholar] [CrossRef]
  114. Matyash, M.; Zabiegalov, O.; Wendt, S.; Matyash, V.; Kettenmann, H. The Adenosine Generating Enzymes CD39/CD73 Control Microglial Processes Ramification in the Mouse Brain. PLoS ONE 2017, 12, e0175012. [Google Scholar] [CrossRef] [PubMed]
  115. Koizumi, S.; Shigemoto-Mogami, Y.; Nasu-Tada, K.; Shinozaki, Y.; Ohsawa, K.; Tsuda, M.; Joshi, B.V.; Jacobson, K.A.; Kohsaka, S.; Inoue, K. UDP Acting at P2Y6 Receptors is a Mediator of Microglial Phagocytosis. Nature 2007, 446, 1091–1095. [Google Scholar] [CrossRef] [PubMed]
  116. Inoue, K.; Tsuda, M. P2X4 Receptors of Microglia in Neuropathic Pain. CNS Neurol. Disord.—Drug Targets 2012, 11, 699–704. [Google Scholar] [CrossRef] [PubMed]
  117. Inoue, K.; Tsuda, M. Microglia in Neuropathic Pain: Cellular and Molecular Mechanisms and Therapeutic Potential. Nat. Rev. Neurosci. 2018, 19, 138–152. [Google Scholar] [CrossRef] [PubMed]
  118. Chessell, I.P.; Hatcher, J.P.; Bountra, C.; Michel, A.D.; Hughes, J.P.; Green, P.; Egerton, J.; Murfin, M.; Richardson, J.; Peck, W.L.; et al. Disruption of the P2X7 Purinoceptor Gene Abolishes Chronic Inflammatory and Neuropathic Pain. Pain 2005, 114, 386–396. [Google Scholar] [CrossRef] [PubMed]
  119. Luchting, B.; Heyn, J.; Woehrle, T.; Rachinger-Adam, B.; Kreth, S.; Hinske, L.C.; Azad, S.C. Differential Expression of P2X7 Receptor and IL-1β in Nociceptive and Neuropathic Pain. J. Neuroinflamm. 2016, 13, 100. [Google Scholar] [CrossRef]
  120. Zhang, W.-J.; Zhu, Z.-M.; Liu, Z.-X. The Role and Pharmacological Properties of the P2X7 Receptor in Neuropathic Pain. Brain Res. Bull. 2020, 155, 19–28. [Google Scholar] [CrossRef]
  121. Di Virgilio, F. Liaisons Dangereuses: P2X7 and the Inflammasome. Trends Pharmacol. Sci. 2007, 28, 465–472. [Google Scholar] [CrossRef]
  122. Tewari, M.; Khan, M.; Verma, M.; Coppens, J.; Kemp, J.M.; Bucholz, R.; Mercier, P.; Egan, T.M. Physiology of Cultured Human Microglia Maintained in a Defined Culture Medium. ImmunoHorizons 2021, 5, 257–272. [Google Scholar] [CrossRef]
  123. Hoshiko, M.; Arnoux, I.; Avignone, E.; Yamamoto, N.; Audinat, E. Deficiency of the Microglial Receptor CX3CR1 Impairs Postnatal Functional Development of Thalamocortical Synapses in the Barrel Cortex. J. Neurosci. 2012, 32, 15106–15111. [Google Scholar] [CrossRef]
  124. Wolf, Y.; Yona, S.; Kim, K.-W.; Jung, S. Microglia, Seen from the CX3CR1 Angle. Front. Cell. Neurosci. 2013, 7, 26. [Google Scholar] [CrossRef] [PubMed]
  125. Swinnen, N.; Smolders, S.; Avila, A.; Notelaers, K.; Paesen, R.; Ameloot, M.; Brône, B.; Legendre, P.; Rigo, J. Complex Invasion Pattern of the Cerebral Cortex Bymicroglial Cells during Development of the Mouse Embryo. Glia 2013, 61, 150–163. [Google Scholar] [CrossRef] [PubMed]
  126. Frost, J.L.; Schafer, D.P. Microglia: Architects of the Developing Nervous System. Trends Cell Biol. 2016, 26, 587–597. [Google Scholar] [CrossRef] [PubMed]
  127. Wolf, S.A.; Boddeke, H.W.G.M.; Kettenmann, H. Microglia in Physiology and Disease. Annu. Rev. Physiol. 2017, 79, 619–643. [Google Scholar] [CrossRef] [PubMed]
  128. Cowan, M.; Petri, W.A. Microglia: Immune Regulators of Neurodevelopment. Front. Immunol. 2018, 9, 2576. [Google Scholar] [CrossRef] [PubMed]
  129. Cunningham, C.L.; Martínez-Cerdeño, V.; Noctor, S.C. Microglia Regulate the Number of Neural Precursor Cells in the Developing Cerebral Cortex. J. Neurosci. 2013, 33, 4216–4233. [Google Scholar] [CrossRef]
  130. Antony, J.M.; Paquin, A.; Nutt, S.L.; Kaplan, D.R.; Miller, F.D. Endogenous Microglia Regulate Development of Embryonic Cortical Precursor Cells. J. Neurosci. Res. 2011, 89, 286–298. [Google Scholar] [CrossRef]
  131. Rymo, S.F.; Gerhardt, H.; Sand, F.W.; Lang, R.; Uv, A.; Betsholtz, C. A Two-Way Communication between Microglial Cells and Angiogenic Sprouts Regulates Angiogenesis in Aortic Ring Cultures. PLoS ONE 2011, 6, e15846. [Google Scholar] [CrossRef]
  132. Reemst, K.; Noctor, S.C.; Lucassen, P.J.; Hol, E.M. The Indispensable Roles of Microglia and Astrocytes during Brain Development. Front. Hum. Neurosci. 2016, 10, 566. [Google Scholar] [CrossRef]
  133. Thion, M.S.; Ginhoux, F.; Garel, S. Microglia and Early Brain Development: An Intimate Journey. Science 2018, 362, 185–189. [Google Scholar] [CrossRef]
  134. Nelson, L.H.; Saulsbery, A.I.; Lenz, K.M. Small Cells with Big Implications: Microglia and Sex Differences in Brain Development, Plasticity and Behavioral Health. Prog. Neurobiol. 2019, 176, 103–119. [Google Scholar] [CrossRef]
  135. Paolicelli, R.C.; Bolasco, G.; Pagani, F.; Maggi, L.; Scianni, M.; Panzanelli, P.; Giustetto, M.; Ferreira, T.A.; Guiducci, E.; Dumas, L.; et al. Synaptic Pruning by Microglia Is Necessary for Normal Brain Development. Science 2011, 333, 1456–1458. [Google Scholar] [CrossRef] [PubMed]
  136. Casano, A.M.; Albert, M.; Peri, F. Developmental Apoptosis Mediates Entry and Positioning of Microglia in the Zebrafish Brain. Cell Rep. 2016, 16, 897–906. [Google Scholar] [CrossRef] [PubMed]
  137. Davalos, D.; Grutzendler, J.; Yang, G.; Kim, J.V.; Zuo, Y.; Jung, S.; Littman, D.R.; Dustin, M.L.; Gan, W.-B. ATP Mediates Rapid Microglial Response to Local Brain Injury In Vivo. Nat. Neurosci. 2005, 8, 752–758. [Google Scholar] [CrossRef] [PubMed]
  138. Kaur, C.; Rathnasamy, G.; Ling, E.-A. Biology of Microglia in the Developing Brain. J. Neuropathol. Exp. Neurol. 2017, 76, 736–753. [Google Scholar] [CrossRef]
  139. Sominsky, L.; De Luca, S.; Spencer, S.J. Microglia: Key Players in Neurodevelopment and Neuronal Plasticity. Int. J. Biochem. Cell Biol. 2018, 94, 56–60. [Google Scholar] [CrossRef]
  140. Cengiz, P.; Zafer, D.; Chandrashekhar, J.H.; Chanana, V.; Bogost, J.; Waldman, A.; Novak, B.; Kintner, D.B.; Ferrazzano, P.A. Developmental Differences in Microglia Morphology and Gene Expression during Normal Brain Development and in Response to Hypoxia-Ischemia. Neurochem. Int. 2019, 127, 137–147. [Google Scholar] [CrossRef]
  141. Nimmerjahn, A.; Kirchhoff, F.; Helmchen, F. Resting Microglial Cells Are Highly Dynamic Surveillants of Brain Parenchyma in Vivo. Science 2005, 308, 1314–1318. [Google Scholar] [CrossRef]
  142. Kierdorf, K.; Prinz, M. Microglia in Steady State. J. Clin. Investig. 2017, 127, 3201–3209. [Google Scholar] [CrossRef]
  143. Madry, C.; Attwell, D. Receptors, Ion Channels, and Signaling Mechanisms Underlying Microglial Dynamics. J. Biol. Chem. 2015, 290, 12443–12450. [Google Scholar] [CrossRef]
  144. Bernier, L.-P.; Bohlen, C.J.; York, E.M.; Choi, H.B.; Kamyabi, A.; Dissing-Olesen, L.; Hefendehl, J.K.; Collins, H.Y.; Stevens, B.; Barres, B.A.; et al. Nanoscale Surveillance of the Brain by Microglia via cAMP-Regulated Filopodia. Cell Rep. 2019, 27, 2895–2908.e4. [Google Scholar] [CrossRef] [PubMed]
  145. Tremblay, M.Ě.; Lowery, R.L.; Majewska, A.K. Microglial Interactions with Synapses Are Modulated by Visual Experience. PLoS Biol. 2010, 8, e1000527. [Google Scholar] [CrossRef]
  146. 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]
  147. Schafer, D.P.; Stevens, B. Microglia Function in Central Nervous System Development and Plasticity. Cold Spring Harb. Perspect. Biol. 2015, 7, a020545. [Google Scholar] [CrossRef] [PubMed]
  148. Cserép, C.; Pósfai, B.; Orsolits, B.; Molnár, G.; Heindl, S.; Lénárt, N.; Fekete, R.; László, Z.I.; Lele, Z.; Schwarcz, A.D.; et al. Microglia Monitor and Protect Neuronal Function Via Specialized Somatic Purinergic Junctions in an Activity-Dependent Manner. SSRN Electron. J. 2019, 367, 528–537. [Google Scholar] [CrossRef]
  149. Hammond, T.R.; Robinton, D.; Stevens, B. Microglia and the Brain: Complementary Partners in Development and Disease. Annu. Rev. Cell Dev. Biol. 2018, 34, 523–544. [Google Scholar] [CrossRef] [PubMed]
  150. Mody, M.; Cao, Y.; Cui, Z.; Tay, K.-Y.; Shyong, A.; Shimizu, E.; Pham, K.; Schultz, P.; Welsh, D.; Tsien, J.Z. Genome-Wide Gene Expression Profiles of the Developing Mouse Hippocampus. Proc. Natl. Acad. Sci. USA 2001, 98, 8862–8867. [Google Scholar] [CrossRef]
  151. Stevens, B.; Allen, N.J.; Vazquez, L.E.; Howell, G.R.; Christopherson, K.S.; Nouri, N.; Micheva, K.D.; Mehalow, A.K.; Huberman, A.D.; Stafford, B.; et al. The Classical Complement Cascade Mediates CNS Synapse Elimination. Cell 2007, 131, 1164–1178. [Google Scholar] [CrossRef]
  152. Györffy, B.A.; Kun, J.; Török, G.; Bulyáki, É.; Borhegyi, Z.; Gulyássy, P.; Kis, V.; Szocsics, P.; Micsonai, A.; Matkó, J.; et al. Local Apoptotic-like Mechanisms Underlie Complement-Mediated Synaptic Pruning. Proc. Natl. Acad. Sci. USA 2018, 115, 6303–6308. [Google Scholar] [CrossRef]
  153. Stephan, A.H.; Barres, B.A.; Stevens, B. The Complement System: An Unexpected Role in Synaptic Pruning during Development and Disease. Annu. Rev. Neurosci. 2012, 35, 369–389. [Google Scholar] [CrossRef]
  154. Thion, M.S.; Garel, S. Microglia Under the Spotlight: Activity and Complement-Dependent Engulfment of Synapses. Trends Neurosci. 2018, 41, 332–334. [Google Scholar] [CrossRef]
  155. Li, W.; Ma, L.; Yang, G.; Gan, W.-B. REM Sleep Selectively Prunes and Maintains New Synapses in Development and Learning. Nat. Neurosci. 2017, 20, 427–437. [Google Scholar] [CrossRef]
  156. Seibt, J.; Frank, M.G. Primed to Sleep: The Dynamics of Synaptic Plasticity Across Brain States. Front. Syst. Neurosci. 2019, 13, 2. [Google Scholar] [CrossRef]
  157. Choudhury, M.E.; Miyanishi, K.; Takeda, H.; Islam, A.; Matsuoka, N.; Kubo, M.; Matsumoto, S.; Kunieda, T.; Nomoto, M.; Yano, H.; et al. Phagocytic Elimination of Synapses by Microglia during Sleep. Glia 2020, 68, 44–59. [Google Scholar] [CrossRef]
  158. Stowell, R.D.; Sipe, G.O.; Dawes, R.P.; Batchelor, H.N.; Lordy, K.A.; Whitelaw, B.S.; Stoessel, M.B.; Bidlack, J.M.; Brown, E.; Sur, M.; et al. Noradrenergic Signaling in the Wakeful State Inhibits Microglial Surveillance and Synaptic Plasticity in the Mouse Visual Cortex. Nat. Neurosci. 2019, 22, 1782–1792. [Google Scholar] [CrossRef]
  159. Wang, C.; Yue, H.; Hu, Z.; Shen, Y.; Ma, J.; Li, J.; Wang, X.-D.; Wang, L.; Sun, B.; Shi, P.; et al. Microglia Mediate Forgetting via Complement-Dependent Synaptic Elimination. Science 2020, 367, 688–694. [Google Scholar] [CrossRef]
  160. Ferrini, F.; De Koninck, Y. Microglia Control Neuronal Network Excitability via BDNF Signalling. Neural Plast. 2013, 2013, 429815. [Google Scholar] [CrossRef]
  161. Pöyhönen, S.; Er, S.; Domanskyi, A.; Airavaara, M. Effects of Neurotrophic Factors in Glial Cells in the Central Nervous System: Expression and Properties in Neurodegeneration and Injury. Front. Physiol. 2019, 10, 486. [Google Scholar] [CrossRef]
  162. Akiyoshi, R.; Wake, H.; Kato, D.; Horiuchi, H.; Ono, R.; Ikegami, A.; Haruwaka, K.; Omori, T.; Tachibana, Y.; Moorhouse, A.J.; et al. Microglia Enhance Synapse Activity to Promote Local Network Synchronization. eNeuro 2018, 5. [Google Scholar] [CrossRef]
  163. Branchi, I.; Alboni, S.; Maggi, L. The Role of Microglia in Mediating the Effect of the Environment in Brain Plasticity and Behavior. Front. Cell. Neurosci. 2014, 8, 390. [Google Scholar] [CrossRef]
  164. Augusto-Oliveira, M.; Arrifano, G.P.; Lopes-Araújo, A.; Santos-Sacramento, L.; Takeda, P.Y.; Anthony, D.C.; Malva, J.O.; Crespo-Lopez, M.E. What Do Microglia Really Do in Healthy Adult Brain? Cells 2019, 8, 1293. [Google Scholar] [CrossRef]
  165. Mariani, M.M.; Kielian, T. Microglia in Infectious Diseases of the Central Nervous System. J. Neuroimmune Pharmacol. 2009, 4, 448–461. [Google Scholar] [CrossRef]
  166. Mogensen, T.H. Pathogen Recognition and Inflammatory Signaling in Innate Immune Defenses. Clin. Microbiol. Rev. 2009, 22, 240–273. [Google Scholar] [CrossRef]
  167. Amarante-Mendes, G.P.; Adjemian, S.; Branco, L.M.; Zanetti, L.C.; Weinlich, R.; Bortoluci, K.R. Pattern Recognition Receptors and the Host Cell Death Molecular Machinery. Front. Immunol. 2018, 9, 2379. [Google Scholar] [CrossRef]
  168. Bsibsi, M.; Ravid, R.; Gveric, D.; van Noort, J.M. Broad Expression of Toll-Like Receptors in the Human Central Nervous System. J. Neuropathol. Exp. Neurol. 2002, 61, 1013–1021. [Google Scholar] [CrossRef]
  169. Fiebich, B.L.; Batista, C.R.A.; Saliba, S.W.; Yousif, N.M.; de Oliveira, A.C.P. Role of Microglia TLRs in Neurodegeneration. Front. Cell. Neurosci. 2018, 12, 329. [Google Scholar] [CrossRef]
  170. Kielian, T. Toll-like Receptors in Central Nervous System Glial Inflammation and Homeostasis. J. Neurosci. Res. 2006, 83, 711–730. [Google Scholar] [CrossRef]
  171. Nie, L.; Cai, S.-Y.; Shao, J.-Z.; Chen, J. Toll-Like Receptors, Associated Biological Roles, and Signaling Networks in Non-Mammals. Front. Immunol. 2018, 9, 1523. [Google Scholar] [CrossRef]
  172. Murao, A.; Aziz, M.; Wang, H.; Brenner, M.; Wang, P. Release Mechanisms of Major DAMPs. Apoptosis 2021, 26, 152–162. [Google Scholar] [CrossRef]
  173. Wright-Jin, E.C.; Gutmann, D.H. Microglia as Dynamic Cellular Mediators of Brain Function. Trends Mol. Med. 2019, 25, 967–979. [Google Scholar] [CrossRef] [PubMed]
  174. Wake, H.; Moorhouse, A.J.; Miyamoto, A.; Nabekura, J. Microglia: Actively Surveying and Shaping Neuronal Circuit Structure and Function. Trends Neurosci. 2013, 36, 209–217. [Google Scholar] [CrossRef] [PubMed]
  175. Song, W.M.; Colonna, M. The Identity and Function of Microglia in Neurodegeneration. Nat. Immunol. 2018, 19, 1048–1058. [Google Scholar] [CrossRef]
  176. Chen, X.; Firulyova, M.; Manis, M.; Herz, J.; Smirnov, I.; Aladyeva, E.; Wang, C.; Bao, X.; Finn, M.B.; Hu, H.; et al. Microglia-Mediated T Cell Infiltration Drives Neurodegeneration in Tauopathy. Nature 2023, 615, 668–677. [Google Scholar] [CrossRef]
  177. Wang, S.; Colonna, M. Microglia in Alzheimer’s Disease: A Target for Immunotherapy. J. Leukoc. Biol. 2019, 106, 219–227. [Google Scholar] [CrossRef]
  178. Deczkowska, A.; Keren-Shaul, H.; Weiner, A.; Colonna, M.; Schwartz, M.; Amit, I. Disease-Associated Microglia: A Universal Immune Sensor of Neurodegeneration. Cell 2018, 173, 1073–1081. [Google Scholar] [CrossRef] [PubMed]
  179. Keren-Shaul, H.; Spinrad, A.; Weiner, A.; Matcovitch-Natan, O.; Dvir-Szternfeld, R.; Ulland, T.K.; David, E.; Baruch, K.; Lara-Astaiso, D.; Toth, B.; et al. A Unique Microglia Type Associated with Restricting Development of Alzheimer’s Disease. Cell 2017, 169, 1276–1290.e17. [Google Scholar] [CrossRef]
  180. Friedman, B.A.; Srinivasan, K.; Ayalon, G.; Meilandt, W.J.; Lin, H.; Huntley, M.A.; Cao, Y.; Lee, S.-H.; Haddick, P.C.; Ngu, H.; et al. Diverse Brain Myeloid Expression Profiles Reveal Distinct Microglial Activation States and Aspects of Alzheimer’s Disease Not Evident in Mouse Models. Cell Rep. 2018, 22, 832–847. [Google Scholar] [CrossRef] [PubMed]
  181. Wang, S.; Sudan, R.; Peng, V.; Zhou, Y.; Du, S.; Yuede, C.M.; Lei, T.; Hou, J.; Cai, Z.; Cella, M.; et al. TREM2 Drives Microglia Response to Amyloid-β via SYK-Dependent and -Independent Pathways. Cell 2022, 185, 4153–4169.e19. [Google Scholar] [CrossRef]
  182. Poliani, P.L.; Wang, Y.; Fontana, E.; Robinette, M.L.; Yamanishi, Y.; Gilfillan, S.; Colonna, M. TREM2 Sustains Microglial Expansion during Aging and Response to Demyelination. J. Clin. Investig. 2015, 125, 2161–2170. [Google Scholar] [CrossRef]
  183. Ulland, T.K.; Colonna, M. TREM2—A Key Player in Microglial Biology and Alzheimer Disease. Nat. Rev. Neurol. 2018, 14, 667–675. [Google Scholar] [CrossRef]
  184. Zhao, Y.; Wu, X.; Li, X.; Jiang, L.-L.; Gui, X.; Liu, Y.; Sun, Y.; Zhu, B.; Piña-Crespo, J.C.; Zhang, M.; et al. TREM2 Is a Receptor for β-Amyloid that Mediates Microglial Function. Neuron 2018, 97, 1023–1031.e7. [Google Scholar] [CrossRef] [PubMed]
  185. Zhong, L.; Wang, Z.; Wang, D.; Wang, Z.; Martens, Y.A.; Wu, L.; Xu, Y.; Wang, K.; Li, J.; Huang, R.; et al. Amyloid-Beta Modulates Microglial Responses by Binding to the Triggering Receptor Expressed on Myeloid Cells 2 (TREM2). Mol. Neurodegener. 2018, 13, 15. [Google Scholar] [CrossRef] [PubMed]
  186. Jain, N.; Lewis, C.A.; Ulrich, J.D.; Holtzman, D.M. Chronic TREM2 Activation Exacerbates Aβ-Associated Tau Seeding and Spreading. J. Exp. Med. 2022, 220, e20220654. [Google Scholar] [CrossRef]
  187. Nacmias, B.; Tedde, A.; Latorraca, S.; Piacentini, S.; Bracco, L.; Amaducci, L.; Guarnieri, B.M.; Petruzzi, C.; Ortenzi, L.; Sorbi, S. Apolipoprotein E and α1-Antichymotrypsin Polymorphism in Alzheimer’s Disease. Ann. Neurol. 1996, 40, 678–680. [Google Scholar] [CrossRef]
  188. Yamazaki, Y.; Zhao, N.; Caulfield, T.R.; Liu, C.-C.; Bu, G. Apolipoprotein E and Alzheimer Disease: Pathobiology and Targeting Strategies. Nat. Rev. Neurol. 2019, 15, 501–518. [Google Scholar] [CrossRef] [PubMed]
  189. Krasemann, S.; Madore, C.; Cialic, R.; Baufeld, C.; Calcagno, N.; El Fatimy, R.; Beckers, L.; O’Loughlin, E.; Xu, Y.; Fanek, Z.; et al. The TREM2-APOE Pathway Drives the Transcriptional Phenotype of Dysfunctional Microglia in Neurodegenerative Diseases. Immunity 2017, 47, 566–581.E9. [Google Scholar] [CrossRef]
  190. Jurga, A.M.; Piotrowska, A.; Makuch, W.; Przewlocka, B.; Mika, J. Blockade of P2X4 Receptors Inhibits Neuropathic Pain-Related Behavior by Preventing MMP-9 Activation and, Consequently, Pronociceptive Interleukin Release in a Rat Model. Front. Pharmacol. 2017, 8, 48. [Google Scholar] [CrossRef]
  191. Di Virgilio, F.; Sarti, A.C. Microglia P2X4 Receptors as Pharmacological Targets for Demyelinating Diseases. EMBO Mol. Med. 2018, 10, e9369. [Google Scholar] [CrossRef]
  192. Zabala, A.; Vazquez-Villoldo, N.; Rissiek, B.; Gejo, J.; Martin, A.; Palomino, A.; Perez-Samartín, A.; Pulagam, K.R.; Lukowiak, M.; Capetillo-Zarate, E.; et al. P2X4 Receptor Controls Microglia Activation and Favors Remyelination in Autoimmune Encephalitis. EMBO Mol. Med. 2018, 10, e8743. [Google Scholar] [CrossRef]
  193. Long, T.; He, W.; Pan, Q.; Zhang, S.; Zhang, Y.; Liu, C.; Liu, Q.; Qin, G.; Chen, L.; Zhou, J. Microglia P2X4 Receptor Contributes to Central Sensitization Following Recurrent Nitroglycerin Stimulation. J. Neuroinflamm. 2018, 15, 245. [Google Scholar] [CrossRef]
  194. Nörenberg, W.; Langosch, J.; Gebicke-Haerter, P.; Illes, P. Characterization and Possible Function of Adenosine 5′-Triphosphate Receptors in Activated Rat Microglia. Br. J. Pharmacol. 1994, 111, 942–950. [Google Scholar] [CrossRef]
  195. Walz, W.; Ilschner, S.; Ohlemeyer, C.; Banati, R.; Kettenmann, H. Extracellular ATP Activates a Cation Conductance and a K+ Conductance in Cultured Microglial Cells from Mouse Brain. J. Neurosci. 1993, 13, 4403–4411. [Google Scholar] [CrossRef] [PubMed]
  196. Illes, P.; Nörenberg, W.; Gebicke-Haerter, P.J. Molecular mechanisms of microglial activation. B. Voltage- and purinoceptor-operated channels in microglia. Neurochem. Int. 1996, 29, 13–24. [Google Scholar] [CrossRef] [PubMed]
  197. Toulme, E.; Garcia, A.; Samways, D.; Egan, T.M.; Carson, M.J.; Khakh, B.S. P2X4 Receptors in Activated C8-B4 Cells of Cerebellar Microglial Origin. J. Gen. Physiol. 2010, 135, 333–353. [Google Scholar] [CrossRef]
  198. 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]
  199. Haas, S.; Brockhaus, J.; Verkhratsky, A.; Kettenmann, H. ATP-Induced Membrane Currents in Ameboid Microglia Acutely Isolated from Mouse Brain Slices. Neuroscience 1996, 75, 257–261. [Google Scholar] [CrossRef]
  200. Schilling, T.; Eder, C. Microglial K+ Channel Expression in Young Adult and Aged Mice. Glia 2015, 63, 664–672. [Google Scholar] [CrossRef] [PubMed]
  201. Nörenberg, W.; Gebicke-Haerter, P.J.; Illes, P. Voltage-Dependent Potassium Channels in Activated Rat Microglia. J. Physiol. 1994, 475, 15–32. [Google Scholar] [CrossRef]
  202. Bianco, F.; Ceruti, S.; Colombo, A.; Fumagalli, M.; Ferrari, D.; Pizzirani, C.; Matteoli, M.; Di Virgilio, F.; Abbracchio, M.P.; Verderio, C. A Role for P2X7 in Microglial Proliferation. J. Neurochem. 2006, 99, 745–758. [Google Scholar] [CrossRef]
  203. Hide, I.; Tanaka, M.; Inoue, A.; Nakajima, K.; Kohsaka, S.; Inoue, K.; Nakata, Y. Extracellular ATP Triggers Tumor Necrosis Factor-α Release from Rat Microglia. J. Neurochem. 2000, 75, 965–972. [Google Scholar] [CrossRef]
  204. Inoue, K.; Nakajima, K.; Morimoto, T.; Kikuchi, Y.; Koizumi, S.; Illes, P.; Kohsaka, S. ATP stimulation of Ca2+-dependent plasminogen release from cultured microglia. Br. J. Pharmacol. 1998, 123, 1304–1310. [Google Scholar] [CrossRef] [PubMed]
  205. Ferrari, D.; Stroh, C.; Schulze-Osthoff, K. P2X7/P2Z Purinoreceptor-mediated Activation of Transcription Factor NFAT in Microglial Cells. J. Biol. Chem. 1999, 274, 13205–13210. [Google Scholar] [CrossRef]
  206. Mackenzie, A.B.; Young, M.T.; Adinolfi, E.; Surprenant, A. Pseudoapoptosis Induced by Brief Activation of ATP-gated P2X7 Receptors. J. Biol. Chem. 2005, 280, 33968–33976. [Google Scholar] [CrossRef] [PubMed]
  207. Parvathenani, L.K.; Tertyshnikova, S.; Greco, C.R.; Roberts, S.B.; Robertson, B.; Posmantur, R. P2X7 Mediates Superoxide Production in Primary Microglia and Is Up-Regulated in a Transgenic Mouse Model of Alzheimer’s Disease. J. Biol. Chem. 2003, 278, 13309–13317. [Google Scholar] [CrossRef]
  208. Pétrilli, V.; Papin, S.; Dostert, C.; Mayor, A.; Martinon, F.; Tschopp, J. Activation of the NALP3 Inflammasome is Triggered by Low Intracellular Potassium Concentration. Cell Death Differ. 2007, 14, 1583–1589. [Google Scholar] [CrossRef]
  209. Swanson, K.V.; Deng, M.; Ting, J.P.-Y. The NLRP3 inflammasome: Molecular Activation and Regulation to Therapeutics. Nat. Rev. Immunol. 2019, 19, 477–489. [Google Scholar] [CrossRef] [PubMed]
  210. Drinkall, S.; Lawrence, C.B.; Ossola, B.; Russell, S.; Bender, C.; Brice, N.B.; Dawson, L.A.; Harte, M.; Brough, D. The Two Pore potassium Channel THIK-1 Regulates NLRP3 Inflammasome Activation. Glia 2022, 70, 1301–1316. [Google Scholar] [CrossRef] [PubMed]
  211. Ossola, B.; Rifat, A.; Rowland, A.; Hunter, H.; Drinkall, S.; Bender, C.; Hamlischer, M.; Teall, M.; Burley, R.; Barker, D.F.; et al. Characterisation of C101248: A Novel Selective THIK-1 Channel Inhibitor for the Modulation of Microglial NLRP3-Inflammasome. Neuropharmacology 2023, 224, 109330. [Google Scholar] [CrossRef]
  212. Di, A.; Xiong, S.; Ye, Z.; Malireddi, R.S.; Kometani, S.; Zhong, M.; Mittal, M.; Hong, Z.; Kanneganti, T.-D.; Rehman, J.; et al. The TWIK2 Potassium Efflux Channel in Macrophages Mediates NLRP3 Inflammasome-Induced Inflammation. Immunity 2018, 49, 56–65.e4. [Google Scholar] [CrossRef]
  213. Chessell, I.P.; Michel, A.D.; Humphrey, P.P.A. Properties of the Pore-Forming P2x7 Purinoceptor in Mouse NTW8 Microglial Cells. Br. J. Pharmacol. 1997, 121, 1429–1437. [Google Scholar] [CrossRef]
  214. Raouf, R.; Chabot-Doré, A.-J.; Ase, A.R.; Blais, D.; Séguéla, P. Differential Regulation of Microglial P2X4 and P2X7 ATP Receptors following LPS-Induced Activation. Neuropharmacology 2007, 53, 496–504. [Google Scholar] [CrossRef]
  215. Steinberg, T.; Silverstein, S. Extracellular ATP4-Promotes Cation Fluxes in the J774 Mouse Macrophage Cell Line. J. Biol. Chem. 1987, 262, 3118–3122. [Google Scholar] [CrossRef] [PubMed]
  216. Franceschini, A.; Capece, M.; Chiozzi, P.; Falzoni, S.; Sanz, J.M.; Sarti, A.C.; Bonora, M.; Pinton, P.; Di Virgilio, F. The P2X7 Receptor Directly Interacts with the NLRP3 Inflammasome Scaffold Protein. FASEB J. 2015, 29, 2450–2461. [Google Scholar] [CrossRef]
  217. Virginio, C.; MacKenzie, A.; North, R.A.; Surprenant, A. Kinetics of Cell Lysis, Dye Uptake and Permeability Changes in Cells Expressing the Rat P2X7 Receptor. J. Physiol. 1999, 519, 335–346. [Google Scholar] [CrossRef]
  218. Di Virgilio, F.; Schmalzing, G.; Markwardt, F. The Elusive P2X7 Macropore. Trends Cell Biol. 2018, 28, 392–404. [Google Scholar] [CrossRef]
  219. Peverini, L.; Beudez, J.; Dunning, K.; Chataigneau, T.; Grutter, T. New Insights Into Permeation of Large Cations Through ATP-Gated P2X Receptors. Front. Mol. Neurosci. 2018, 11, 265. [Google Scholar] [CrossRef] [PubMed]
  220. Rozengurt, E.; Heppel, L.A. A Specific Effect of External ATP on the Permeability of Transformed 3T3 Cells. Biochem. Biophys. Res. Commun. 1975, 67, 1581–1588. [Google Scholar] [CrossRef] [PubMed]
  221. Cockcroft, S.; Gomperts, B.D. ATP Induces Nucleotide Permeability in Rat Mast Cells. Nature 1979, 279, 541–542. [Google Scholar] [CrossRef]
  222. Steinberg, T.H.; Newman, A.S.; Swanson, J.A.; Silverstein, S.C. ATP4-Permeabilizes the Plasma Membrane of Mouse Macrophages to Fluorescent Dyes. J. Biol. Chem. 1987, 262, 8884–8888. [Google Scholar] [CrossRef]
  223. Buisman, H.P.; Steinberg, T.H.; Fischbarg, J.; Silverstein, S.C.; Vogelzang, S.A.; Ince, C.; Ypey, D.L.; Leijh, P.C. Extracellular ATP Induces a Large Nonselective Conductance in Macrophage Plasma Membranes. Proc. Natl. Acad. Sci. USA 1988, 85, 7988–7992. [Google Scholar] [CrossRef]
  224. Ferrari, D.; Villalba, M.; Chiozzi, P.; Falzoni, S.; Ricciardi-Castagnoli, P.; Di Virgilio, F. Mouse Microglial Cells Express a Plasma Membrane Pore Gated by Extracellular ATP. J. Immunol. 1996, 156, 1531–1539. [Google Scholar] [CrossRef]
  225. Khakh, B.S.; Bao, X.R.; Labarca, C.; Lester, H.A. Neuronal P2X Transmitter-Gated Cation Channels Change Their Ion Selectivity in Seconds. Nat. Neurosci. 1999, 2, 322–330. [Google Scholar] [CrossRef] [PubMed]
  226. Virginio, C.; MacKenzie, A.; Rassendren, F.A.; North, R.A.; Surprenant, A. Pore Dilation of Neuronal P2X Receptor Channels. Nat. Neurosci. 1999, 2, 315–321. [Google Scholar] [CrossRef] [PubMed]
  227. Harkat, M.; Peverini, L.; Cerdan, A.H.; Dunning, K.; Beudez, J.; Martz, A.; Calimet, N.; Specht, A.; Cecchini, M.; Chataigneau, T.; et al. On the Permeation of Large Organic Cations through the Pore of ATP-Gated P2X Receptors. Proc. Natl. Acad. Sci. USA 2017, 114, E3786–E3795. [Google Scholar] [CrossRef]
  228. Chung, M.-K.; Güler, A.D.; Caterina, M.J. TRPV1 Shows Dynamic Ionic Selectivity during Agonist Stimulation. Nat. Neurosci. 2008, 11, 555–564. [Google Scholar] [CrossRef] [PubMed]
  229. Karasawa, A.; Michalski, K.; Mikhelzon, P.; Kawate, T. The P2X7 Receptor Forms a Dye-Permeable Pore Independent of Its Intracellular Domain but Dependent on Membrane Lipid Composition. eLife 2017, 6, e31186. [Google Scholar] [CrossRef]
  230. Pelegrin, P.; Surprenant, A. Pannexin-1 Mediates Large Pore Formation and Interleukin-1β Release by the ATP-Gated P2X7 Receptor. EMBO J. 2006, 25, 5071–5082. [Google Scholar] [CrossRef] [PubMed]
  231. Pelegrin, P.; Surprenant, A. The P2X7 Receptor–Pannexin Connection to Dye Uptake and IL-1β Release. Purinergic Signal. 2009, 5, 129–137. [Google Scholar] [CrossRef]
  232. Janks, L.; Sprague, R.S.; Egan, T.M. ATP-Gated P2X7 Receptors Require Chloride Channels to Promote Inflammation in Human Macrophages. J. Immunol. 2019, 202, 883–898. [Google Scholar] [CrossRef]
  233. Duan, S.; Anderson, C.M.; Keung, E.C.; Chen, Y.; Chen, Y.; Swanson, R.A. P2X7 Receptor-Mediated Release of Excitatory Amino Acids from Astrocytes. J. Neurosci. 2003, 23, 1320–1328. [Google Scholar] [CrossRef]
  234. Marques-Da-Silva, C.; Chaves, M.M.; Rodrigues, J.C.; Corte-Real, S.; Coutinho-Silva, R.; Persechini, P.M. Differential Modulation of ATP-Induced P2X7-Associated Permeabilities to Cations and Anions of Macrophages by Infection with Leishmania amazonensis. PLoS ONE 2011, 6, e25356. [Google Scholar] [CrossRef]
  235. Ugur, M.; Ugur, Ö. A Mechanism-Based Approach to P2X7 Receptor Action. Mol. Pharmacol. 2019, 95, 442–450. [Google Scholar] [CrossRef]
  236. Schachter, J.; Motta, A.P.; Zamorano, A.d.S.; da Silva-Souza, H.A.; Guimarães, M.Z.P.; Persechini, P.M. ATP-Induced P2X7-Associated Uptake of Large Molecules Involves Distinct Mechanisms for Cations and Anions in Macrophages. J. Cell Sci. 2008, 121, 3261–3270. [Google Scholar] [CrossRef]
  237. Ferrari, D.; Chiozzi, P.; Falzoni, S.; Susino, M.D.; Collo, G.; Buell, G.; Di Virgilio, F. ATP-Mediated Cytotoxicity in Microglial Cells. Neuropharmacology 1997, 36, 1295–1301. [Google Scholar] [CrossRef] [PubMed]
  238. Guerra, G.P.; Rubin, M.A.; Mello, C.F. Modulation of Learning and Memory by Natural Polyamines. Pharmacol. Res. 2016, 112, 99–118. [Google Scholar] [CrossRef]
  239. Charras, G.T.; Coughlin, M.; Mitchison, T.J.; Mahadevan, L. Life and Times of a Cellular Bleb. Biophys. J. 2008, 94, 1836–1853. [Google Scholar] [CrossRef] [PubMed]
  240. Weng, N.J.-H.; Talbot, P. The P2X7 Receptor is an Upstream Regulator of Dynamic Blebbing and a Pluripotency Marker in Human Embryonic Stem Cells. Stem Cell Res. 2017, 23, 39–49. [Google Scholar] [CrossRef] [PubMed]
  241. MacKenzie, A.; Wilson, H.L.; Kiss-Toth, E.; Dower, S.K.; North, R.A.; Surprenant, A. Rapid Secretion of Interleukin-1β by Microvesicle Shedding. Immunity 2001, 15, 825–835. [Google Scholar] [CrossRef]
  242. Bianco, F.; Perrotta, C.; Novellino, L.; Francolini, M.; Riganti, L.; Menna, E.; Saglietti, L.; Schuchman, E.H.; Furlan, R.; Clementi, E.; et al. Acid Sphingomyelinase Activity Triggers Microparticle Release from Glial Cells. EMBO J. 2009, 28, 1043–1054. [Google Scholar] [CrossRef]
  243. Charras, G.; Paluch, E. Blebs Lead the Way: How to Migrate without Lamellipodia. Nat. Rev. Mol. Cell Biol. 2008, 9, 730–736. [Google Scholar] [CrossRef]
  244. Babiychuk, E.B.; Monastyrskaya, K.; Potez, S.; Draeger, A. Blebbing Confers Resistance against Cell Lysis. Cell Death Differ. 2011, 18, 80–89. [Google Scholar] [CrossRef]
  245. Di Virgilio, F.; Dal Ben, D.; Sarti, A.C.; Giuliani, A.L.; Falzoni, S. The P2X7 Receptor in Infection and INFLAMMATION. Immunity 2017, 47, 15–31. [Google Scholar] [CrossRef]
  246. Inoue, K. Microglial Activation by Purines and Pyrimidines. Glia 2002, 40, 156–163. [Google Scholar] [CrossRef] [PubMed]
  247. Shieh, C.; Heinrich, A.; Serchov, T.; van Calker, D.; Biber, K. P2X7-Dependent, but Differentially Regulated Release of IL-6, CCL2, and TNF-α in Cultured Mouse Microglia. Glia 2014, 62, 592–607. [Google Scholar] [CrossRef] [PubMed]
  248. Savio, L.E.B.; de Andrade Mello, P.; Da Silva, C.G.; Coutinho-Silva, R. The P2X7 Receptor in Inflammatory Diseases: Angel or Demon? Front. Pharmacol. 2018, 9, 52. [Google Scholar] [CrossRef] [PubMed]
  249. Sanz, J.M.; Di Virgilio, F. Kinetics and Mechanism of ATP-Dependent IL-1β Release from Microglial Cells. J. Immunol. 2000, 164, 4893–4898. [Google Scholar] [CrossRef] [PubMed]
  250. Suzuki, T.; Hide, I.; Ido, K.; Kohsaka, S.; Inoue, K.; Nakata, Y. Production and Release of Neuroprotective Tumor Necrosis Factor by P2X7 Receptor-Activated Microglia. J. Neurosci. 2004, 24, 1–7. [Google Scholar] [CrossRef]
  251. Facci, L.; Barbierato, M.; Zusso, M.; Skaper, S.D.; Giusti, P. Serum Amyloid A Primes Microglia for ATP-Dependent Interleukin-1β Release. J. Neuroinflamm. 2018, 15, 164. [Google Scholar] [CrossRef]
  252. Ferrari, D.; Pizzirani, C.; Adinolfi, E.; Lemoli, R.M.; Curti, A.; Idzko, M.; Panther, E.; Di Virgilio, F. The P2X7 Receptor: A Key Player in IL-1 Processing and Release. J. Immunol. 2006, 176, 3877–3883. [Google Scholar] [CrossRef]
  253. Munoz, F.M.; Patel, P.A.; Gao, X.; Mei, Y.; Xia, J.; Gilels, S.; Hu, H. Reactive Oxygen Species Play a Role in P2X7 Receptor-Mediated IL-6 Production in Spinal Astrocytes. Purinergic Signal. 2020, 16, 97–107. [Google Scholar] [CrossRef]
  254. Kim, S.Y.; Moon, J.H.; Lee, H.G.; Kim, S.U.; Lee, Y.B. ATP Released from β-Amyloid-Stimulated Microglia Induces Reactive Oxygen Species Production in an Autocrine Fashion. Exp. Mol. Med. 2007, 39, 820–827. [Google Scholar] [CrossRef]
  255. Quail, D.F.; Joyce, J.A. Microenvironmental Regulation of Tumor Progression and Metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef] [PubMed]
  256. Adinolfi, E.; De Marchi, E.; Orioli, E.; Pegoraro, A.; Di Virgilio, F. Role of the P2X7 Receptor in Tumor-Associated Inflammation. Curr. Opin. Pharmacol. 2019, 47, 59–64. [Google Scholar] [CrossRef] [PubMed]
  257. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, Inflammation, and Cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef]
  258. Amoroso, F.; Capece, M.; Rotondo, A.; Cangelosi, D.; Ferracin, M.; Franceschini, A.; Raffaghello, L.; Pistoia, V.; Varesio, L.; Adinolfi, E. The P2X7 Receptor is a Key Modulator of the PI3K/GSK3β/VEGF Signaling Network: Evidence in Experimental Neuroblastoma. Oncogene 2015, 34, 5240–5251. [Google Scholar] [CrossRef] [PubMed]
  259. Di Virgilio, F.; Adinolfi, E. Extracellular Purines, Purinergic Receptors and Tumor Growth. Oncogene 2017, 36, 293–303. [Google Scholar] [CrossRef]
  260. Scarpellino, G.; Genova, T.; Munaron, L. Purinergic P2X7 Receptor: A Cation Channel Sensitive to Tumor Microenvironment. Recent Patents Anti-Cancer Drug Discov. 2019, 14, 32–38. [Google Scholar] [CrossRef] [PubMed]
  261. Arnaud-Sampaio, V.F.; Rabelo, I.L.A.; Ulrich, H.; Lameu, C. The P2X7 Receptor in the Maintenance of Cancer Stem Cells, Chemoresistance and Metastasis. Stem Cell Rev. Rep. 2019, 16, 288–300. [Google Scholar] [CrossRef]
  262. Hui, L.; Chen, Y. Tumor Microenvironment: Sanctuary of the Devil. Cancer Lett. 2015, 368, 7–13. [Google Scholar] [CrossRef]
  263. Bianchi, G.; Vuerich, M.; Pellegatti, P.; Marimpietri, D.; Emionite, L.; Marigo, I.; Bronte, V.; Di Virgilio, F.; Pistoia, V.; Raffaghello, L. ATP/P2X7 Axis Modulates Myeloid-Derived Suppressor Cell Functions in Neuroblastoma Microenvironment. Cell Death Dis. 2014, 5, e1135. [Google Scholar] [CrossRef]
  264. Adinolfi, E.; Raffaghello, L.; Giuliani, A.L.; Cavazzini, L.; Capece, M.; Chiozzi, P.; Bianchi, G.; Kroemer, G.; Pistoia, V.; Di Virgilio, F. Expression of P2X7 Receptor Increases In Vivo Tumor Growth. Cancer Res 2012, 72, 2957–2969. [Google Scholar] [CrossRef] [PubMed]
  265. Bergamin, L.S.; Capece, M.; Salaro, E.; Sarti, A.C.; Falzoni, S.; Pereira, M.S.L.; De Bastiani, M.A.; Scholl, J.N.; Battastini, A.M.O.; Di Virgilio, F. Role of the P2X7 Receptor in In Vitro and In Vivo Glioma Tumor Growth. Oncotarget 2019, 10, 4840–4856. [Google Scholar] [CrossRef]
  266. Virgilio, F.D. Purines, Purinergic Receptors, and Cancer. Cancer Res. 2012, 72, 5441–5447. [Google Scholar] [CrossRef] [PubMed]
  267. Zhang, Y.; Li, F.; Wang, L.; Lou, Y. A438079 Affects Colorectal Cancer Cell Proliferation, migration, apoptosis, and pyroptosis by inhibiting the P2X7 receptor. Biochem. Biophys. Res. Commun. 2021, 558, 147–153. [Google Scholar] [CrossRef]
  268. Kan, L.K.; Williams, D.; Drummond, K.; O’Brien, T.; Monif, M. The Role of Microglia and P2X7 Receptors in Gliomas. J. Neuroimmunol. 2019, 332, 138–146. [Google Scholar] [CrossRef]
  269. Ghiringhelli, F.; Apetoh, L.; Tesniere, A.; Aymeric, L.; Ma, Y.; Ortiz, C.; Vermaelen, K.; Panaretakis, T.; Mignot, G.; Ullrich, E.; et al. Activation of the NLRP3 inflammasome in Dendritic Cells Induces IL-1β–Dependent Adaptive Immunity against Tumors. Nat. Med. 2009, 15, 1170–1178. [Google Scholar] [CrossRef] [PubMed]
  270. Adinolfi, E.; Capece, M.; Franceschini, A.; Falzoni, S.; Giuliani, A.L.; Rotondo, A.; Sarti, A.C.; Bonora, M.; Syberg, S.; Corigliano, D.; et al. Accelerated Tumor Progression in Mice Lacking the ATP Receptor P2X7. Cancer Res 2015, 75, 635–644. [Google Scholar] [CrossRef]
  271. Hoelzinger, D.B.; Demuth, T.; Berens, M.E. Autocrine Factors That Sustain Glioma Invasion and Paracrine Biology in the Brain Microenvironment. JNCI J. Natl. Cancer Inst. 2007, 99, 1583–1593. [Google Scholar] [CrossRef]
  272. Cekic, C.; Day, Y.-J.; Sag, D.; Linden, J. Myeloid Expression of Adenosine A2A Receptor Suppresses T and NK Cell Responses in the Solid Tumor Microenvironment. Cancer Res 2014, 74, 7250–7259. [Google Scholar] [CrossRef]
  273. Tarassishin, L.; Lim, J.; Weatherly, D.B.; Angeletti, R.H.; Lee, S.C. Interleukin-1-Induced Changes in the Glioblastoma Secretome Suggest Its Role in Tumor Progression. J. Proteom. 2014, 99, 152–168. [Google Scholar] [CrossRef]
  274. Illes, P. P2X7 Receptors Amplify CNS Damage in Neurodegenerative Diseases. Int. J. Mol. Sci. 2020, 21, 5996. [Google Scholar] [CrossRef]
  275. Tewari, M.; Seth, P. Emerging Role of P2X7 Receptors in CNS Health and Disease. Ageing Res. Rev. 2015, 24, 328–342. [Google Scholar] [CrossRef]
  276. Hogquist, K.A.; Nett, M.A.; Unanue, E.R.; Chaplin, D.D. Interleukin 1 is Processed and Released during Apoptosis. Proc. Natl. Acad. Sci. USA 1991, 88, 8485–8489. [Google Scholar] [CrossRef]
  277. Brough, D.; Le Feuvre, R.A.; Iwakura, Y.; Rothwell, N.J. Purinergic (P2X7) Receptor Activation of Microglia Induces Cell Death via an Interleukin-1-Independent Mechanism. Mol. Cell. Neurosci. 2002, 19, 272–280. [Google Scholar] [CrossRef]
  278. Mehta, V.B.; Hart, J.; Wewers, M.D. ATP-stimulated Release of Interleukin (IL)-1β and IL-18 Requires Priming by Lipopolysaccharide and Is Independent of Caspase-1 Cleavage. J. Biol. Chem. 2001, 276, 3820–3826. [Google Scholar] [CrossRef] [PubMed]
  279. Denlinger, L.C.; Fisette, P.L.; Sommer, J.A.; Watters, J.J.; Prabhu, U.; Dubyak, G.R.; Proctor, R.A.; Bertics, P.J. Cutting Edge: The Nucleotide Receptor P2X7 Contains Multiple Protein- and Lipid-Interaction Motifs Including a Potential Binding Site for Bacterial Lipopolysaccharide. J. Immunol. 2001, 167, 1871–1876. [Google Scholar] [CrossRef] [PubMed]
  280. Bartlett, R.; Yerbury, J.J.; Sluyter, R. P2X7 Receptor Activation Induces Reactive Oxygen Species Formation and Cell Death in Murine EOC13 Microglia. Mediat. Inflamm. 2013, 2013, 271813. [Google Scholar] [CrossRef] [PubMed]
  281. Loureiro, A.V.; Moura-Neto, L.I.; Martins, C.S.; Silva, P.I.M.; Lopes, M.B.; Leitão, R.F.C.; Coelho-Aguiar, J.M.; Moura-Neto, V.; Warren, C.A.; Costa, D.V.; et al. Role of Pannexin-1-P2X7R Signaling on Cell Death and Pro-Inflammatory Mediator Expression Induced by Clostridioides Difficile Toxins in Enteric Glia. Front. Immunol. 2022, 13, 956340. [Google Scholar] [CrossRef]
  282. Cavaliere, F.; Dinkel, K.; Reymann, K. Microglia Response and P2 Receptor Participation in Oxygen/Glucose Deprivation-Induced Cortical Damage. Neuroscience 2005, 136, 615–623. [Google Scholar] [CrossRef]
  283. Cavaliere, F.; D’Ambrosi, N.; Sancesario, G.; Bernardi, G.; Volonté, C. Hypoglycaemia-Induced Cell Death: Features of Neuroprotection by the P2 Receptor Antagonist Basilen Blue. Neurochem. Int. 2000, 38, 199–207. [Google Scholar] [CrossRef]
  284. Eyo, U.B.; Miner, S.A.; Ahlers, K.E.; Wu, L.-J.; Dailey, M.E. P2X7 Receptor Activation Regulates Microglial Cell Death during Oxygen-Glucose Deprivation. Neuropharmacology 2013, 73, 311–319. [Google Scholar] [CrossRef] [PubMed]
  285. Jia, X.; Xie, L.; Liu, Y.; Liu, T.; Yang, P.; Hu, J.; Peng, Z.; Luo, K.; Du, M.; Chen, C. Astragalus polysaccharide (APS) Exerts Protective Effect against Acute Ischemic Stroke (AIS) through Enhancing M2 Micoglia Polarization by Regulating Adenosine Triphosphate (ATP)/Purinergic Receptor (P2X7R) Axis. Bioengineered 2022, 13, 4468–4480. [Google Scholar] [CrossRef]
  286. Lopez, J.A.S.; González, H.M.; Léger, G.C. Alzheimer’s Disease. In Handbook of Clinical Neurology; Elsevier: Amsterdam, The Netherlands, 2019; Volume 167, pp. 231–255. [Google Scholar]
  287. De-Paula, V.J.; Radanovic, M.; Diniz, B.S.; Forlenza, O.V. Alzheimer’s Disease. Subcell. Biochem. 2012, 65, 329–352. [Google Scholar]
  288. Wang, W.-Y.; Tan, M.-S.; Yu, J.-T.; Tan, L. Role of Pro-Inflammatory Cytokines Released from Microglia in Alzheimer’s Disease. Ann. Transl. Med. 2015, 3, 136. [Google Scholar]
  289. Nizami, S.; Hall-Roberts, H.; Warrier, S.; Cowley, S.A.; Di Daniel, E. Microglial Inflammation and Phagocytosis in Alzheimer’s Disease: Potential Therapeutic Targets. Br. J. Pharmacol. 2019, 176, 3515–3532. [Google Scholar] [CrossRef] [PubMed]
  290. Suresh, P.; Phasuk, S.; Liu, I. Modulation of Microglia Activation and Alzheimer’s Disease: CX3 Chemokine Ligand 1/CX3CR and P2X7R Signaling. Tzu Chi Med. J. 2021, 33, 1–6. [Google Scholar] [CrossRef] [PubMed]
  291. McLarnon, J.G.; Ryu, J.K.; Walker, D.G.; Choi, H.B. Upregulated Expression of Purinergic P2X(7) Receptor in Alzheimer Disease and Amyloid-Beta Peptide-Treated Microglia and in Peptide-Injected Rat Hippocampus. J. Neuropathol. Exp. Neurol. 2006, 65, 1090–1097. [Google Scholar] [CrossRef]
  292. Lee, H.G.; Won, S.M.; Gwag, B.J.; Lee, Y.B. Microglial P2X Receptor Expression Is Accompanied by Neuronal Damage in the Cerebral Cortex of the APPswe/PS1dE9 Mouse Model of Alzheimer’s Disease. Exp. Mol. Med. 2011, 43, 7–14. [Google Scholar] [CrossRef]
  293. 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]
  294. 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]
  295. Thawkar, B.S.; Kaur, G. Inhibitors of NF-κB and P2X7/NLRP3/Caspase 1 Pathway in Microglia: Novel Therapeutic Opportunities in Neuroinflammation Induced Early-Stage Alzheimer’s Disease. J. Neuroimmunol. 2019, 326, 62–74. [Google Scholar] [CrossRef]
  296. Francistiová, L.; Bianchi, C.; Di Lauro, C.; Sebastián-Serrano, Á.; De Diego-García, L.; Kobolák, J.; Dinnyés, A.; Díaz-Hernández, M. The Role of P2X7 Receptor in Alzheimer’s Disease. Front. Mol. Neurosci. 2020, 13, 94. [Google Scholar] [CrossRef]
  297. Chen, Y.-H.; Lin, R.-R.; Tao, Q.-Q. The role of P2X7R in Neuroinflammation and Implications in Alzheimer’s Disease. Life Sci. 2021, 271, 119187. [Google Scholar] [CrossRef]
  298. Carvalho, K.; Martin, E.; Ces, A.; Sarrazin, N.; Lagouge-Roussey, P.; Nous, C.; Boucherit, L.; Youssef, I.; Prigent, A.; Faivre, E.; et al. P2X7-Deficiency Improves Plasticity and Cognitive Abilities in a Mouse Model of Tauopathy. Prog. Neurobiol. 2021, 206, 102139. [Google Scholar] [CrossRef]
  299. Jeon, S.G.; Kang, M.; Kim, Y.-S.; Kim, D.-H.; Nam, D.W.; Song, E.J.; Mook-Jung, I.; Moon, M. Intrahippocampal Injection of a Lentiviral Vector Expressing Neurogranin Enhances Cognitive Function in 5XFAD Mice. Exp. Mol. Med. 2018, 50, e461. [Google Scholar] [CrossRef] [PubMed]
  300. Martin, E.; Amar, M.; Dalle, C.; Youssef, I.; Boucher, C.; Le Duigou, C.; Brückner, M.; Prigent, A.; Sazdovitch, V.; Halle, A.; et al. New Role of P2X7 Receptor in an Alzheimer’s disease Mouse Model. Mol. Psychiatry 2019, 24, 108–125. [Google Scholar] [CrossRef] [PubMed]
  301. Martínez-Frailes, C.; Di Lauro, C.; Bianchi, C.; de Diego-García, L.; Sebastián-Serrano, Á.; Boscá, L.; Díaz-Hernández, M. Amyloid Peptide Induced Neuroinflammation Increases the P2X7 Receptor Expression in Microglial Cells, Impacting on Its Functionality. Front. Cell. Neurosci. 2019, 13, 143. [Google Scholar] [CrossRef] [PubMed]
  302. Cardoso, S.M.; Santana, I.; Swerdlow, R.H.; Oliveira, C.R. Mitochondria Dysfunction of Alzheimer’s Disease Cybrids Enhances Aβ Toxicity. J. Neurochem. 2004, 89, 1417–1426. [Google Scholar] [CrossRef]
  303. Chen, J.X.; Yan, S.D. Amyloid-β-Induced Mitochondrial Dysfunction. J. Alzheimer’s Dis. 2007, 12, 177–184. [Google Scholar] [CrossRef]
  304. Tobore, T.O. On the Central Role of Mitochondria Dysfunction and Oxidative Stress in Alzheimer’s Disease. Neurol. Sci. 2019, 40, 1527–1540. [Google Scholar] [CrossRef]
  305. Sanz, J.M.; Falzoni, S.; Rizzo, R.; Cipollone, F.; Zuliani, G.; Di Virgilio, F. Possible Protective Role of the 489C>T P2X7R Polymorphism in Alzheimer’s Disease. Exp. Gerontol. 2014, 60, 117–119. [Google Scholar] [CrossRef]
  306. Islam, J.; Cho, J.-A.; Kim, J.-Y.; Park, K.-S.; Koh, Y.-J.; Chung, C.Y.; Lee, E.-J.; Nam, S.J.; Lee, K.; Kim, S.-H.; et al. GPCR19 Regulates P2X7R-Mediated NLRP3 Inflammasomal Activation of Microglia by Amyloid β in a Mouse Model of Alzheimer’s Disease. Front. Immunol. 2022, 13, 766919. [Google Scholar] [CrossRef] [PubMed]
  307. Yu, Y.; Jiang, X.; Fang, X.; Wang, Y.; Liu, P.; Ling, J.; Yu, L.; Jiang, M.; Tang, C. Transauricular Vagal Nerve Stimulation at 40 Hz Inhibits Hippocampal P2X7R/NLRP3/Caspase-1 Signaling and Improves Spatial Learning and Memory in 6-Month-Old APP/PS1 Mice. Neuromodul. Technol. Neural Interface 2023, 26, 589–600. [Google Scholar] [CrossRef]
  308. Doorn, K.J.; Moors, T.; Drukarch, B.; van de Berg, W.D.; Lucassen, P.J.; van Dam, A.-M. Microglial Phenotypes and Toll-like Receptor 2 in the Substantia Nigra and Hippocampus of Incidental Lewy Body Disease Cases and Parkinson’s Disease Patients. Acta Neuropathol. Commun. 2014, 2, 90. [Google Scholar] [CrossRef] [PubMed]
  309. Joers, V.; Tansey, M.G.; Mulas, G.; Carta, A.R. Microglial phenotypes in Parkinson’s disease and animal models of the disease. Prog. Neurobiol. 2017, 155, 57–75. [Google Scholar] [CrossRef] [PubMed]
  310. Carmo, M.R.; Menezes, A.P.F.; Nunes, A.C.L.; Pliássova, A.; Rolo, A.P.; Palmeira, C.M.; Cunha, R.A.; Canas, P.M.; Andrade, G.M. The P2X7 receptor antagonist Brilliant Blue G attenuates contralateral rotations in a rat model of Parkinsonism through a combined control of synaptotoxicity, neurotoxicity and gliosis. Neuropharmacology 2014, 81, 142–152. [Google Scholar] [CrossRef] [PubMed]
  311. Van Weehaeghe, D.; Koole, M.; Schmidt, M.E.; Deman, S.; Jacobs, A.H.; Souche, E.; Serdons, K.; Sunaert, S.; Bormans, G.; Vandenberghe, W.; et al. [11C]JNJ54173717, a Novel P2X7 Receptor Radioligand as Marker for Neuroinflammation: Human Biodistribution, Dosimetry, Brain Kinetic Modelling and Quantification of Brain P2X7 Receptors in Patients with Parkinson’s Disease and Healthy Volunteers. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 2051–2064. [Google Scholar] [CrossRef]
  312. Ren, C.; Li, L.-X.; Dong, A.-Q.; Zhang, Y.-T.; Hu, H.; Mao, C.-J.; Wang, F.; Liu, C.-F. Depression Induced by Chronic Unpredictable Mild Stress Increases Susceptibility to Parkinson’s Disease in Mice via Neuroinflammation Mediated by P2X7 Receptor. ACS Chem. Neurosci. 2021, 12, 1262–1272. [Google Scholar] [CrossRef]
  313. Jiang, T.; Hoekstra, J.; Heng, X.; Kang, W.; Ding, J.; Liu, J.; Chen, S.; Zhang, J. P2X7 Receptor is Critical in α-Synuclein–Mediated Microglial NADPH Oxidase Activation. Neurobiol. Aging 2015, 36, 2304–2318. [Google Scholar] [CrossRef]
  314. Crabbé, M.; Van der Perren, A.; Bollaerts, I.; Kounelis, S.; Baekelandt, V.; Bormans, G.; Casteels, C.; Moons, L.; Van Laere, K. Increased P2X7 Receptor Binding Is Associated with Neuroinflammation in Acute but Not Chronic Rodent Models for Parkinson’s Disease. Front. Neurosci. 2019, 13, 799. [Google Scholar] [CrossRef]
  315. Kumar, S.; Mishra, A.; Krishnamurthy, S. Purinergic Antagonism Prevents Mitochondrial Dysfunction and Behavioral Deficits Associated with Dopaminergic Toxicity Induced by 6-OHDA in Rats. Neurochem. Res. 2017, 42, 3414–3430. [Google Scholar] [CrossRef] [PubMed]
  316. Wang, X.-H.; Xie, X.; Luo, X.-G.; Shang, H.; He, Z.-Y. Inhibiting Purinergic P2X7 Receptors with the Antagonist Brilliant Blue G is Neuroprotective in an Intranigral Lipopolysaccharide Animal Model of Parkinson’s Disease. Mol. Med. Rep. 2017, 15, 768–776. [Google Scholar] [CrossRef] [PubMed]
  317. Jiang, T.; Xu, C.; Gao, S.; Zhang, J.; Zheng, J.; Wu, X.; Lu, Q.; Cao, L.; Yang, D.; Xu, J.; et al. Cathepsin L-Containing Exosomes from α-Synuclein-Activated Microglia Induce Neurotoxicity through the P2X7 Receptor. npj Park. Dis. 2022, 8, 127. [Google Scholar] [CrossRef] [PubMed]
  318. Yang, X.; Lou, Y.; Liu, G.; Wang, X.; Qian, Y.; Ding, J.; Chen, S.; Xiao, Q. Microglia P2Y6 Receptor Is Related to Parkinson’s Disease through Neuroinflammatory Process. J. Neuroinflamm. 2017, 14, 38. [Google Scholar] [CrossRef]
  319. Oliveira-Giacomelli, Á.; Albino, C.M.; de Souza, H.D.N.; Corrêa-Velloso, J.; Santos, A.P.d.J.; Baranova, J.; Ulrich, H. P2Y6 and P2X7 Receptor Antagonism Exerts Neuroprotective/Neuroregenerative Effects in an Animal Model of Parkinson’s Disease. Front. Cell. Neurosci. 2019, 13, 476. [Google Scholar] [CrossRef] [PubMed]
  320. Thijs, R.D.; Surges, R.; O’Brien, T.J.; Sander, J.W. Epilepsy in Adults. Lancet 2019, 393, 689–701. [Google Scholar] [CrossRef] [PubMed]
  321. Fisher, R.S.; Acevedo, C.; Arzimanoglou, A.; Bogacz, A.; Cross, J.H.; Elger, C.E.; Engel, J., Jr.; Forsgren, L.; French, J.A.; Glynn, M.; et al. ILAE Official Report: A Practical Clinical Definition of Epilepsy. Epilepsia 2014, 55, 475–482. [Google Scholar] [CrossRef]
  322. Boison, D. Adenosine Dysfunction and Adenosine Kinase in Epileptogenesis. Open Neurosci. J. 2010, 4, 93–101. [Google Scholar] [CrossRef]
  323. Eyo, U.B.; Murugan, M.; Wu, L. Microglia–Neuron Communication in Epilepsy. Glia 2017, 65, 5–18. [Google Scholar] [CrossRef]
  324. Chen, F.; He, X.; Luan, G.; Li, T. Role of DNA Methylation and Adenosine in Ketogenic Diet for Pharmacoresistant Epilepsy: Focus on Epileptogenesis and Associated Comorbidities. Front. Neurol. 2019, 10, 119. [Google Scholar] [CrossRef]
  325. Alyu, F.; Dikmen, M. Inflammatory Aspects of epileptogenesis: Contribution of Molecular Inflammatory Mechanisms. Acta Neuropsychiatr. 2017, 29, 1–16. [Google Scholar] [CrossRef] [PubMed]
  326. Abiega, O.; Beccari, S.; Diaz-Aparicio, I.; Nadjar, A.; Layé, S.; Leyrolle, Q.; Gómez-Nicola, D.; Domercq, M.; Pérez-Samartín, A.; Sánchez-Zafra, V.; et al. Neuronal Hyperactivity Disturbs ATP Microgradients, Impairs Microglial Motility, and Reduces Phagocytic Receptor Expression Triggering Apoptosis/Microglial Phagocytosis Uncoupling. PLOS Biol. 2016, 14, e1002466. [Google Scholar] [CrossRef] [PubMed]
  327. Engel, T.; Alves, M.; Sheedy, C.; Henshall, D.C. ATPergic Signalling during Seizures and Epilepsy. Neuropharmacology 2016, 104, 140–153. [Google Scholar] [CrossRef] [PubMed]
  328. Wang, M.; Deng, X.; Xie, Y.; Chen, Y. Astaxanthin Attenuates Neuroinflammation in Status Epilepticus Rats by Regulating the ATP-P2X7R Signal. Drug Des. Dev. Ther. 2020, 14, 1651–1662. [Google Scholar] [CrossRef] [PubMed]
  329. Dale, N.; Frenguelli, B.G. Release of Adenosine and ATP during Ischemia and Epilepsy. Curr. Neuropharmacol. 2009, 7, 160–179. [Google Scholar] [CrossRef] [PubMed]
  330. Weissberg, I.; Reichert, A.; Heinemann, U.; Friedman, A. Blood-Brain Barrier Dysfunction in Epileptogenesis of the Temporal Lobe. Epilepsy Res. Treat. 2011, 2011, 143908. [Google Scholar] [CrossRef]
  331. Kim, S.Y.; Buckwalter, M.; Soreq, H.; Vezzani, A.; Kaufer, D. Blood-Brain Barrier Dysfunction-Induced Inflammatory Signaling in Brain Pathology and Epileptogenesis. Epilepsia 2012, 53 (Suppl. S6), 37–44. [Google Scholar] [CrossRef]
  332. Hong, S.; Xin, Y.; JiaWen, W.; ShuQin, Z.; GuiLian, Z.; HaiQin, W.; Zhen, G.; HongWei, R.; YongNan, L. The P2X7 Receptor in Activated Microglia Promotes Depression- and Anxiety-like Behaviors in Lithium -Pilocarpine Induced Epileptic Rats. Neurochem. Int. 2020, 138, 104773. [Google Scholar] [CrossRef]
  333. Lee, D.-S.; Kim, J.-E. Protein Disulfide Isomerase-Mediated S-Nitrosylation Facilitates Surface Expression of P2X7 Receptor Following Status Epilepticus. J. Neuroinflamm. 2021, 18, 14. [Google Scholar] [CrossRef]
  334. Zhang, X.; Wang, M.; Feng, B.; Zhang, Q.; Tong, J.; Wang, M.; Lu, C.; Peng, S. Seizures in PPT1 Knock-In Mice Are Associated with Inflammatory Activation of Microglia. Int. J. Mol. Sci. 2022, 23, 5586. [Google Scholar] [CrossRef]
  335. Smith, J.; Méndez, A.M.; Alves, M.; Parras, A.; Conte, G.; Bhattacharya, A.; Ceusters, M.; Nicke, A.; Henshall, D.C.; Jimenez-Mateos, E.M.; et al. The P2X7 Receptor Contributes to Seizures and Inflammation-Driven Long-Lasting Brain Hyperexcitability Following Hypoxia in Neonatal Mice. Br. J. Pharmacol. 2023, 180, 1710–1729. [Google Scholar] [CrossRef] [PubMed]
Figure 1. P2X signaling in microglia. Extracellular ATP evokes influx of Na+ and Ca2+ and efflux of K+ through P2X7Rs embedded in the plasmalemmal membrane. ATP-gated efflux of K+ is also thought to involve additional channels; at present, the identity of these channels is unknown but may involve 2–pore K+ channels. The decrease in [K+]i results in influx of Cl and the maturation and release of the proinflammatory cytokine IL–1β. Activation of P2X7Rs also increases membrane permeabilization of large organic cations like YO-PRO1, decreases phagocytosis, and initiates phosphatidylserine (PS) exposure on the cell surface. While YO-PRO1 permeates the P2X7R pore, other pathways (marked with a “?”) are also thought to play a role.
Figure 1. P2X signaling in microglia. Extracellular ATP evokes influx of Na+ and Ca2+ and efflux of K+ through P2X7Rs embedded in the plasmalemmal membrane. ATP-gated efflux of K+ is also thought to involve additional channels; at present, the identity of these channels is unknown but may involve 2–pore K+ channels. The decrease in [K+]i results in influx of Cl and the maturation and release of the proinflammatory cytokine IL–1β. Activation of P2X7Rs also increases membrane permeabilization of large organic cations like YO-PRO1, decreases phagocytosis, and initiates phosphatidylserine (PS) exposure on the cell surface. While YO-PRO1 permeates the P2X7R pore, other pathways (marked with a “?”) are also thought to play a role.
Cells 13 00161 g001
Figure 2. Model for microglial-driven neuroinflammation. Neuroinflammation can be subdivided into acute and chronic phases. Acute neuroinflammation is characterized by activation of resident immune cell populations, release of proinflammatory mediators including cytokines/chemokines, infiltration of peripheral immune cells to assist in the immune response, and phagocytic activity to reduce cellular debris. When inflammation persists past the initial injury event and homeostasis is not maintained, chronic neuroinflammation commences. This phase is characterized by sustained microglial activation and gliosis measured by heightened release of proinflammatory cytokines/chemokines, reduced phagocytosis, and neurodegeneration. Created with BioRender.com accessed on 8 January 2024.
Figure 2. Model for microglial-driven neuroinflammation. Neuroinflammation can be subdivided into acute and chronic phases. Acute neuroinflammation is characterized by activation of resident immune cell populations, release of proinflammatory mediators including cytokines/chemokines, infiltration of peripheral immune cells to assist in the immune response, and phagocytic activity to reduce cellular debris. When inflammation persists past the initial injury event and homeostasis is not maintained, chronic neuroinflammation commences. This phase is characterized by sustained microglial activation and gliosis measured by heightened release of proinflammatory cytokines/chemokines, reduced phagocytosis, and neurodegeneration. Created with BioRender.com accessed on 8 January 2024.
Cells 13 00161 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tewari, M.; Michalski, S.; Egan, T.M. Modulation of Microglial Function by ATP-Gated P2X7 Receptors: Studies in Rat, Mice and Human. Cells 2024, 13, 161. https://doi.org/10.3390/cells13020161

AMA Style

Tewari M, Michalski S, Egan TM. Modulation of Microglial Function by ATP-Gated P2X7 Receptors: Studies in Rat, Mice and Human. Cells. 2024; 13(2):161. https://doi.org/10.3390/cells13020161

Chicago/Turabian Style

Tewari, Manju, Stephanie Michalski, and Terrance M. Egan. 2024. "Modulation of Microglial Function by ATP-Gated P2X7 Receptors: Studies in Rat, Mice and Human" Cells 13, no. 2: 161. https://doi.org/10.3390/cells13020161

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