Next Article in Journal
Epigallocatechin-3-Gallate Prevents the Acquisition of a Cancer Stem Cell Phenotype in Ovarian Cancer Tumorspheres through the Inhibition of Src/JAK/STAT3 Signaling
Next Article in Special Issue
Efficacy and Safety of [177Lu]Lu-DOTA-TATE in Adults with Inoperable or Metastatic Somatostatin Receptor-Positive Pheochromocytomas/Paragangliomas, Bronchial and Unknown Origin Neuroendocrine Tumors, and Medullary Thyroid Carcinoma: A Systematic Literature Review
Previous Article in Journal
The Potential of SGLT-2 Inhibitors in the Treatment of Polycystic Ovary Syndrome: The Current Status and Future Perspectives
Previous Article in Special Issue
A Single Arm Clinical Study on the Effects of Continuous Erythropoietin Receptor Activator Treatment in Non-Dialysis Patients with Chronic Heart Failure and Renal Anemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

NLRP3 Inflammasome’s Activation in Acute and Chronic Brain Diseases—An Update on Pathogenetic Mechanisms and Therapeutic Perspectives with Respect to Other Inflammasomes

1
Human Histology & Embryology Section, Department of Surgery, Dentistry, Pediatrics, and Gynecology, University of Verona, 37134 Verona, Italy
2
Department of Neurology, Southwest Hospital, Chongqing 400038, China
*
Authors to whom correspondence should be addressed.
Biomedicines 2023, 11(4), 999; https://doi.org/10.3390/biomedicines11040999
Submission received: 7 February 2023 / Revised: 16 March 2023 / Accepted: 17 March 2023 / Published: 23 March 2023

Abstract

:
Increasingly prevalent acute and chronic human brain diseases are scourges for the elderly. Besides the lack of therapies, these ailments share a neuroinflammation that is triggered/sustained by different innate immunity-related protein oligomers called inflammasomes. Relevant neuroinflammation players such as microglia/monocytes typically exhibit a strong NLRP3 inflammasome activation. Hence the idea that NLRP3 suppression might solve neurodegenerative ailments. Here we review the recent Literature about this topic. First, we update conditions and mechanisms, including RNAs, extracellular vesicles/exosomes, endogenous compounds, and ethnic/pharmacological agents/extracts regulating NLRP3 function. Second, we pinpoint NLRP3-activating mechanisms and known NLRP3 inhibition effects in acute (ischemia, stroke, hemorrhage), chronic (Alzheimer’s disease, Parkinson’s disease, Huntington’s disease, MS, ALS), and virus-induced (Zika, SARS-CoV-2, and others) human brain diseases. The available data show that (i) disease-specific divergent mechanisms activate the (mainly animal) brains NLRP3; (ii) no evidence proves that NLRP3 inhibition modifies human brain diseases (yet ad hoc trials are ongoing); and (iii) no findings exclude that concurrently activated other-than-NLRP3 inflammasomes might functionally replace the inhibited NLRP3. Finally, we highlight that among the causes of the persistent lack of therapies are the species difference problem in disease models and a preference for symptomatic over etiologic therapeutic approaches. Therefore, we posit that human neural cell-based disease models could drive etiological, pathogenetic, and therapeutic advances, including NLRP3’s and other inflammasomes’ regulation, while minimizing failure risks in candidate drug trials.

Graphical Abstract

1. Introduction

1.1. An Overall Picture

Acute and chronic human brain diseases have been attracting the increased attention of scientists and the public. This has been due to the concurrence of several factors, i.e., brain illnesses’ mounting prevalence, the persistent lack of effective therapies, increasingly huge healthcare and economic costs, hardships in assisting such patients particularly at home, marked psychopathological impacts on patients and relatives, a greater sensitivity to improper lifestyle consequences, and a common aspiration to long-lasting and healthy aging. To this must be added the growing concern about the serious risk that severe acute brain injuries surreptitiously evolve into chronic neuropathologies such as Alzheimer’s disease (AD), Parkinson’s disease (PD), and amyotrophic lateral sclerosis (ALS). Worldwide yearly estimates of acute brain injuries total about 42 million cases, while symptomatic AD by itself affects more than 50 million people. It is predicted that such figures will double or treble in twenty/thirty years unless effective therapies become available [1,2]. Yet, the latter quite understandable wish is hampered by ongoing controversies due to the still unclarified underlying pathogenetic mechanisms. A common feature in all brain diseases is ongoing neuroinflammation. From this observation, the hypothesis has been put forward that this inflammation is a main causative factor, whose mitigation or suppression would slow or stop the progression and/or improve the outcome [3,4].
“Inflammation” is a physiological defensive reaction of living tissues to harm, aiming at ridding the causative factor(s), disposing of cell debris, and restoring tissue integrity and homeostasis in the short term. In his treatise “De Medicina”, Roman physician Aulus Cornelius Celsius (~14–37 AD; [5]) first described acute inflammation’s five cardinal symptoms, i.e., “rubor” (Lat. reddening) and “calor” (Lat. heat), due to local increases in blood flow; “tumor” (Lat. swelling) caused by edema and leukocyte infiltration due to altered vessel permeability; “dolor” (Lat. pain), elicited by local acidosis overstimulating the nerves; and “laesa functio” (Lat. impaired function”), the injury’s downstream upshot. Conversely, a persisting (chronic) inflammation is a pathological condition whose upshot can be severe.
Obviously, neuroinflammation has specific features, particularly in the various neurodegenerative diseases. In the latter, its onset can be early (familial cases) or surreptitious (sporadic cases). Its course is often quite slow, so that it can progress undetected for decades. However, while unnoticed, chronic neuroinflammation spreads from the site of origin (e.g., frontotemporal cerebral cortex, hippocampus, locus coeruleus, spinal cord) to other regions and in so doing progressively destroys the brain’s neuronal functional reserve. When the reserve is depleted, the gray matter of the cerebral cortex, basal ganglia, thalamus, brain stem, cerebellum, spinal cord, and the white matter connectome (axons) are remarkably thinned. At this stage, the diseases become symptomatic. Progressive decreases in abilities, such as memory, cognition, emotions, psychic, and motor activities, render the patients unable to cope. Eventually, the neuropathology inexorably and more rapidly moves toward the obitus [6,7]. The etiologic factors also trigger various collateral cellular processes, such as the overproduction of hydroxyl radicals, superoxide anions (reactive oxygen species or ROS), nitric oxide (NO), peroxynitrite, ionic dyshomeostasis, mitochondrial, lysosomal, and autophagy disfunctions, and overproduction and accumulation of toxic protein species, which sustain the neuroinflammation. Other events concur, such as leukocyte infiltration and alterations in blood–brain barrier (BBB) function. Altogether, such noxae drive positive feedback loops, aggravating the neuropathology [8,9,10,11,12,13].
Since Celsius’s time, and particularly in the last century, a huge amount of knowledge has been accumulating about the crucial relation between inflammation’s drivers and immunity. Nowadays, we know that the innate immune system secures the first protection against harmful factors or “molecular patterns”. The endogenous damage-associated molecular patterns (DAMPs) and homeostasis-altering molecular patterns (HAMPs) are sterile compounds (e.g., ATP, mitochondrial DNA), dysfunctional metabolism products, and cell debris. The exogenous pathogen-associated molecular patterns (PAMPs) are infectious (bacteria, fungi, viruses, prions) or toxic agents (chemicals, organic molecules). DAMPs/HAMPs/PAMPs form complexes with multiligand cellular “pattern recognition receptors” (PRRs). In turn, such complexes nucleate the assembly of multicomponent protein platforms, the “inflammasomes” [4,14], the activated signaling of which drives the tissue inflammation at the injury’s site.

NLRs Assembly and Signaling Activation

The PRRs’ group names are based upon shared structural domains. The most noted PRRs comprise the NLRs (NOD-like nucleotide-binding domain and leucine-rich-repeat (LRR) family of receptors); ALRs (absent in melanoma 2 receptors); and MEFV gene-encoded PYRIN receptors [15]. Currently, activated NLRs are the most intensely studied PRRs. In humans, NLRs having a PYRIN N-terminal homology domain (PYD) include 14 members, namely, NLRP1–NLR14. Physiologically, NLRs (excepting brain NLRs) keep an auto-inhibited conformation that winds up when they detect DAMPs/PAMPs/HAMPs. This drives the assembly and signaling activation of inflammasomes. NLRs’ N-terminal PYDs bind and nucleate the oligomerizing adaptor protein ASC (apoptosis-associated speck-like protein endowed with a caspase recruitment domain or CARD) [15,16]. Notably, the ASC gene encodes both a CARD and a PYD domain. Therefore, via CARD•CARD or PYD•PYD homotypic interactions, ASC proteins make complexes with the PYD or CARD domains of NLRs. PYDs and CARDs are conserved domains of 80–90 amino acids arranged in six anti-parallel α-helices forming an inner hydrophobic core with charged residues at the surface. Via CARD•CARD interactions, ASCs of canonical inflammasomes nucleate the inactive zymogens of caspase-1, a cysteine-type peptidase, causing their polymerization and proximity-mediated auto-catalytic self-cleavage, resulting in active caspase-1 duplets [16,17]. The latter produce mature interleukin (IL)-1β and IL-18 from their respective precursors and N-termini fragments of the gasdermin D protein (human, GSDMD; rodent, GsdmD), in addition to cleaving other proteins that share the YVHD/FESD consensus sequence [18]. Next, the GSDMD/GsdmD’s N-fragments oligomerize, forming transmembrane pores that extracellularly release (i) mature proinflammatory IL-1β and IL-18; and (ii) K+, causing an intracellular ion dyshomeostasis. Persistent K+ losses lead to inflammatory death or pyroptosis of the involved cells. In turn, products released from pyroptotic cells (e.g., ATP, mitochondrial DNA) boost inflammation further [18]. NLRP oligomerization, ASC recruitment, and caspase-1 nucleated polymerization/activation are irreversible processes developing in a self-inducing prion-like fashion and promoting canonical inflammasome signaling [19].
Moreover, via CARD, domain-assembled NLRP1, NLRP2, NLRP3, and AIM2 inflammasomes activate the NF-κB signaling pathway, which transcriptionally regulates the genes encoding for the various inflammasomes’ structural proteins [20]. Conversely, other NLRs, i.e., NLRC3, NLRP6, NLRP12, and NLRX1, impede the NF-κB pathway’s activation, thereby mitigating or quelling inflammation [21]. Indeed, these “anti-inflammasomes” are crucially necessary, as they stop the onset of chronic inflammatory diseases. Moreover, CARD-only proteins (COPs) and PYD-only proteins (POPs) also regulate inflammasome activity [22]. Furthermore, epigenetic mechanisms, e.g., noncoding RNA expression, CpG island DNA methylation, and histone post-translational changes, modulate inflammasome function [23].
We recently reviewed the multiple roles of the NLRP1, NLRP2, AIM-2, and NLRC4 inflammasomes in human and rodent brain diseases [24]. Our work showed that several inflammasomes can partake in brain neuroinflammatory processes. This enticed us to review in this work the mounting literature specifically concerning the NLRP3 inflammasome, its modulation by endogenous and exogenous and pharmacological and ethnopharmacological agents/extracts, its pathogenetic implications in acute and chronic brain diseases, and the therapeutic potential of its inhibition. Based on the results we highlight that disease-specific divergent mechanisms activate the brain’s NLRP3 in microglia/monocytes and other neural cell types. However, no proof is hitherto available that NLRP3 inhibition would be a human brain disease-modifying approach. Furthermore, no data have so far excluded the possible functional replacement of the inhibited NLRP3 by other concurrently activated inflammasomes. These facts led us to highlight that one of the causes of the persisting failures of human brain disease-related therapeutic attempts is the inadequate regard for its morpho-functional uniqueness based on the assumption that animal brain models are good enough. The consequent suggestion is to focus instead on human neural cell-based preclinical brain diseases models, which could drive etiological, pathogenetic, and therapeutic advances, including proper NLRP3 and other inflammasome regulation, and minimize failure risks concerning lead candidate drug testing in clinical trials.
The following paragraphs will delve into the main advances concerning the NLRP3 inflammasome, followed by specific paragraphs about its role in most relevant brain diseases, a discussion of the results, and a conclusion.

1.2. Brain NLRP3 Inflammasome

The inactive NLRP3 inflammasome (i.e., NLRP3-ASC or NOD-like receptor protein 3 (N-terminal PYD, central ATP-hydrolyzing NACHT (NAIP+CIITA+HET-E+TP1), and C-terminal LRR domains) molecules confine themselves to the endoplasmic reticulum (ER) membranes [25]. Upon activation, they bind adaptor ASC proteins by interacting with phosphatidylinositol-4-phosphate. ASC stabilizes the NLRP3•ASC complexes allowing their activation. Next, NLRP3•ASC complexes migrate to the perinuclear ER membranes and associated mitochondrial aggregates [9,26].
As monocytes/macrophages and microglia strongly express the NLRP3 inflammasome, the latter is involved in human brain diseases and is the most intensely studied and popular inflammasome. NLRP3 might be the “golden” therapeutic target of inflammatory morbidities, including neurodegenerative disorders (e.g., Alzheimer’s disease [AD]) [27,28,29]. In advanced age, the NLRP3 inflammasome also partakes in low-grade sterile yet chronic inflammation called “inflammaging”, driven by cell debris accumulating within tissues [30]. Moreover, NLRP3 gene mutations result in a spectrum of autoinflammatory diseases known as cryopyrin-associated periodic syndromes (CAPS) [31].
Table 1 lists the common brain NLRP3 inflammasome-activating diseases or agents.

1.2.1. NLRP3 Inflammasome Priming and Canonical Activation

Importantly, human, and rodent brain cells of all types preferentially express distinct inflammasomes, e.g., NLRP1 the neurons, NLRP2 the astrocytes, and NLRP3 the microglia [123,124,125,126,127,128,129]. However, under both normal and pathological conditions, all the neural cell types express the NLRP3 inflammasome, albeit with differing intensities and regulatory mechanisms [27,64]. Young mice brains physiologically express basal levels of NLRP3 inflammasome activity to upkeep conditioning-induced neuronal plasticity and memory consolidation in the ventral hippocampus and basolateral amygdala [130]. Discordant opinions exist about inflammasomes’ roles in human brain diseases, as specific molecular lines of evidence are scanty [24,131,132].
Most studies have shown that NLRP3′s canonical activation requires two initiating signals. The “Signal 1” or “priming step” is an endocytosed PAMP or an endogenous DAMP/HAMP evoking the signaling from Toll-like receptor 4 (TLR-4) or a NOD-like receptor (NLR) or the tumor necrosis factor receptor (TNFR). Furthermore, signaling from G-protein-coupled receptors (GPCRs) can affect NLRP3 activity (see Box 1 for further details and references).
Box 1. NLRP3 inflammasome regulation by G-protein coupled receptors (GPCRs).
The six GPCRs families (A–F) include eight hundred entities. The fact that 34% of FDA-approved drugs target GPCRs proves their clinical importance. For space reasons here, we discuss only a few GPCRs. For further information, see [133].
  B1.1. Calcium-Sensing Receptor (CaSR)
  The extracellular domain (i.e., venus flytrap) of the ubiquitously expressed CaSR of family C GPCRs binds not only Ca2+, its orthosteric (type I) agonist, but also other mono-, bi-, and tri-valent cations, and various positively charged organic molecules, including polyamines, aminoglycoside antibiotics, and cationic peptides (e.g., amyloid-β [Aβ]) [134,135,136]. Moreover, CaSR’s 7TM (seven-pass transmembrane domain) binds allosteric (type II) ligands (e.g., aromatic L-α-amino acids) and positive allosteric modulators (PAMs i.e., calcimimetics) and negative allosteric modulators (NAMs i.e., calcilytics). Ligand-activated CaSR signaling by its intracellular domains is mediated by various G-proteins and scaffold proteins (e.g., β-arrestin, homer-1) and turns on or off several pathways involving various enzymes, ion channels, and transcription factors [133]. Acting as a calciostat sensing changes in [Ca2+]e, the CaSR regulates systemic [Ca2+]e homeostasis via parathormone secretion, modulating gut Ca2+ absorption, bone Ca2+ storage/release, and renal Ca2+ excretion [137]. All types of neural cells express the CaSR, and those in AD-relevant hippocampus very intensely [138]. Importantly, besides [Ca2+]e homeostasis, the CaSR physiologically regulates neural cell growth, differentiation, migration, synaptic plasticity, and neurotransmission [133]. Moreover, the CaSR acts as a DAMP/HAMP/PAMP sensor, as inflammatory diseases affecting various organs, brain included, activate CaSR signaling [27]. In turn, CaSR signaling activates the NLRP3 inflammasome via a surge in phospholipase C-mediated [Ca2+]i and a concurrent fall in the NLRP3-inhibiting cAMP [31], as well as a proteolytic cleavage of crucial NLRP3 regulators [139]. Moreover, increasing cAMP levels via an adenylate cyclase (AC) activator (e.g., PGE2) or a covalently changed (e.g., dibutyryl-) cAMP or a phosphodiesterase (PDE) inhibitor blocking cAMP catabolism to 5′-AMP (e.g., theophylline or milrinone) promotes cAMP binding to NLRP3, which hinders its activation [26,31,140]. CaSR PAM cinacalcet activates NLRP3 inflammasome via ERK1/2 signaling [98]. Wang et al. [99] showed that in subarachnoid hemorrhage-model mice, CaSR’s expression surged in all CNS cell types. The CaSR agonist gadolinium trichloride (GdCI3) upregulated the levels of phosphorylated CaMKII, NLRP3 inflammasome expression, active caspase-1, and mature IL-1β. Conversely, CaSR NAM NPS-2143 and CAMKII inhibitor KN-93 mitigated all CaSR signaling detrimental effects. Hence, CaSR signaling advanced the first stages of acute brain injury, and Aβ•CaSR signaling could drive human AD onset/progression [141].
  B1.2. G-Protein-Coupled Class C Group 6 Receptor A (GPC6RA)
  Alum has been and still is in use as an adjuvant in human vaccines. Alum’s mechanism of action remained obscure until Quandt et al. [50] proved that in vitro and in vivo alum induced NLRP3 inflammasome activation via GPRC6A receptor signaling. GPC6RA, of the GPCR Family C Group 6, senses cations (e.g., Ca2+), osteocalcin, L-α-amino acids, and testosterone. GPC6RA signaling partakes via MAPK and mTORC1 in prostatic carcinoma progression [51,142,143,144,145] and might contribute to the angiotensin II-driven hypertensive neuroinflammation promoted by 6β-hydroxytestosterone in male mice [146].
  B1.3. G protein-coupled estrogen receptors (GPERs)
  GPER1 and GPER30 are seven-pass transmembrane orphan receptors that rapidly mediate non-genomic estrogen-related kinase signaling. GPER signals prevented hippocampal neuron death due to transient global cerebral ischemia via a remarkable elevation of the endogenous interleukin-1 receptor antagonist (IL-1Ra), which suppresses the pro-inflammatory effects of IL-1β. GPER activation heightened the hippocampal levels of phosphorylated CREB (i.e., cAMP response element-binding) transcription factor, which promotes IL-1Ra expression. The G36 antagonist reversed GPER’s neuroprotective effects, proving their specificity [147].
  Clearly, CaSR, GPC6RA, and GPERs are PRRs whose roles in neuroinflammation are worthy of further investigation.
Signal 1 involves both translational and post-translational pathways linked to IFNR, PKA, MAPK, mTOR, complement proteins, AMPK/autophagy, IRAK1, TRIF (TIR[Toll/IL-1 receptor/resistance protein]-domain-containing adapter-inducing IFN-β), and NLRP3’s de-ubiquination by BRCC3 (BRCA1/BRCA2-Containing Complex Subunit 3), a Lys63-specific de-ubiquitinase. These pathways converge toward NF-κB pathway’s activation, which mediates the genetic transcription of NLRP3, ASC, pro-caspase-1, pro-IL-1β, and pro-IL-18 [148,149,150,151,152]. The contours of “Signal 2” or the “activation step” of the NLRP3 inflammasome are less defined. A summary list of Signal 2 includes exogenous dead cell-released ATP, which is a ligand of purinergic receptors (see Box 2 for further details and references); cathepsin B released from destabilized lysosomes; phagocytosed protein polymers; reactive oxygen species (ROS); cardiolipin; oxidized mitochondrial DNA [112,114]; K+ efflux or Ca2+ influx, independently of each other [153]; and cyclic AMP (cAMP) downregulation [154]. Importantly, also contact sites between mitochondria and ER membranes favor NLRP3 activation. ER-stress signal-released mitochondrial proteins, ER-released Ca2+ surges, lipid perturbations, and cholesterol trafficking critically partake in NLRP3 activation [155]. Moreover, a surge in extracellular Ca2+ ([Ca2+]e) triggers NLRP3 activation in monocytes [156]. Thus, [Ca2+]i increases might be the signal shared by all the stimuli [155] and/or the final common NLRP3-activating pathway [157,158].
Box 2. Brain purinergic receptors.
CNS neural cells express diverse types of purinergic receptors, i.e., P1, for adenosine G protein-coupled receptors; P2X, for ATP-gated ion channels; and P2Y, for G protein-coupled receptors. Importantly, the intra-brain accumulation of Aβs induces the damaged neural cells to release ATP into the extracellular matrix (ECM). Exogenous ATP and the agonist 4-benzoyl-ATP (BzATP) activate the signaling from P2X7 purinergic receptors expressed by neural cells. The upshots are an increased synthesis and release of pro-inflammatory cytokines and chemokines, and a decline in the α-secretase activity, causing a plunge in the extracellular shedding of the neurotrophic and neuroprotective soluble amyloid precursor protein (APP)-α. Yet, various (e.g., mechanical) stressing factors awaken the signaling of P2X7 receptors, making the cells release their endogenous ATP through connexin 43 and pannexin hemichannels (i.e., “pathological pores”) [159]. The results are the activation of the NF-κB axis and of the NLRP3•ASC•caspase-1 and IL-1β pathways in both the astrocytes and microglia, triggering the sterile neuroinflammation proper of AD within the brain and of glaucoma within the retina [57,160].
  Moreover, the P2X7 receptor agonist BzATP also elicits the release of various cytokines from the retinal ganglion neurons, i.e., IL-3 (in the presence of extracellular Ca2+); IL-4; IL-10; IL-1Ra; TNF-α; MIG/CXCL9 (or monokine induced by IFN-γ/chemokine [C–X–C motif] ligand 9); VEGF; GM-CSF; MIP (macrophage inflammatory protein); CCL20 (or chemokine [C–C motif] ligand 20); and L-selectin, which altogether exert neuroprotective effects [161]. P2X7 receptor stimulation also upregulates IL-6 release from the retinal astrocytes and neurons [162]. In microglial cells, P2X7 receptors modulate the phagocytosis of exogenous debris in the absence of any ligand. However, signals from ligand-bound P2X7 alter lysosome function, causing the cathepsin B-mediated NLRP3 inflammasome activation that a cathepsin B-blocker, CA-074, instead hinders [163].
  P2X7 −/− (KO), P2X7 antagonists, such as Brilliant Blue G (BBG), A438079, A839977 and A740003, and the NF-κB inhibitor Bay 11-7082 blocked the effects elicited by purinergic receptors signaling. However, P2X7-specific antagonists blocked only the purinergic receptor-dependent secretion of IL-6 and CCL2 but not TNF-α’s release from microglia. These results revealed the differential regulation of the microglial secretion of such cytokines [164]. By contrast, the ATP-activated signaling from the P2Y2 purinergic receptor exerted P2X7-opposite, i.e., anti-inflammatory, and neuroprotective effects [165,166].
Nuclear receptors too control the NLRP3 inflammasome [167]. Thus, various positive and negative signaling pathways strictly regulate NLRP3’s activation to prevent any harm while preserving the host tissues’ homeostasis [168]. Various kinases, ubiquitin ligases, a de-ubiquitinase, and other enzymes crucially control both NLRP3’s activation and function termination via ad hoc post-translational modifications of its protein components [169]. As an example, Bruton’s tyrosine kinase (BTK) directly and positively regulates the NLRP3 inflammasome, which might have therapeutic implications [170]. Usually, sterile, and slow-acting DAMPs/HAMPs elicit weaker NLRP3 inflammasome responses than infectious PAMPS do [171]. Finally, inflammasome-interested scientists should note that species-related differences in animal models can crucially affect their results [172].

1.2.2. Noncanonical NLRP3 Activation

Hitherto, we have discussed NLRP3’s “canonical activation”, a concept valid also for NLRP1, NLRC4, and AIM2 inflammasomes. The more recently discovered “noncanonical activation” of inflammasomes is worth mentioning too. Concerning microglia’s NLRP3, the noncanonical process involves the activation of caspase-11 and caspase-8 in mice and of caspase-4 and caspase-5 in humans [173,174,175]. These caspases behave as cytosolic sensors that directly bind and are activated by the lipopolysaccharide (LPS) of Gram-negative bacteria. This drives the secretion of mature IL-1β and IL-18. Additionally, the active caspases detach N-terminal fragments from the GSDMD/GsdmD proteins, which form transmembrane pores promoting K+ efflux and thus causing both NLRP3’s canonical activation and neurons’ pyroptosis [176,177,178,179].
The HMGB1 (high mobility group box 1 protein)/caspase-8 pathway is an added mechanism of noncanonical NLRP3 activation proper of eye glaucoma. An acutely elevated intraocular pressure intensifies HMGB1’s signaling, which activates the NLRP3 inflammasome by canonical and noncanonical (via caspase-8) mechanisms, producing higher amounts of mature IL-1β within the ischemic retinal tissue and thereby advancing neuroinflammation [59].

1.3. Brain NLRP3 Inflammasome’s Modulation by RNAs

Cells express manifold kinds (ribosomal, messenger, and noncoding) of RNAs, which control most of their functions. Long noncoding (Lnc) RNAs have more than 200 base pairs but encode no or few proteins. However, LncRNAs importantly affect body development, cell differentiation, metabolism, autoimmunity, and immune function, and hence NLRP3 inflammasome activity [180,181]. MicroRNAs (or miRs) are ubiquitous 22-nucleotide-long single-stranded RNAs that post-transcriptionally control gene expression by silencing mRNAs via complementary base-pairing [182]. Notably, miRs abound (>2300 types) inside mammalian cells and are released via extracellular vesicles (EVs) or exosomes (Exos) into cerebrospinal fluid and blood. Circulating miRs are under investigation as biomarkers in various diseases and in the distinct stages of each illness. According to ongoing circumstances, distinct miRs promote or inhibit NLRP3 inflammasome activation.
Among noncoding RNAs, Alu-derived RNAs deserve a brief mention. They result from the transcription of primate-specific transposable “Alu elements” by small interspersed nuclear elements (SINEs). Alu-RNAs are plentiful, involving >10% of the human genome, with 102 to 103 copies released into the cytosol of each cell. Alu-RNAs regulate gene expression by binding and inhibiting RNA polymerase II (P2). Alu-RNAs accumulate in the brains of patients with dementia or sporadic Creutzfeldt–Jacob’s disease (CJD), in which they drive neuroinflammation and neuron demise [183]. P3-transcribed Alu-RNAs (P3Alus) may advance NLRP3 inflammasome-driven neuroinflammation/neurodegeneration disorders, AD included [184]. Hence P3Alus may be therapeutic targets for such ailments. Later studies revealed that Alu-RNAs processing rates are elevated in mouse and human AD brains, tightly correlating with the up-regulated expression of HSF1 (heat shock transcription factor 1), a crucial stress response factor. The increased Alu-RNAs processing rates would fix into active mode the HSF1/Alu-RNA/stress response/cell death-promoting genes (e.g., p53) axis in AD patients [185,186].
This topic is bound to undergo further developments in regard not only to LncRNAs, miRs, and Alu-RNAs, but also to the recently discovered circular RNAs [187].
Table 2 reports details about LncRNAs/miRs and NLRP3 interactions.

1.4. Brain NLRP3 Inflammasome’s Modulation by Extracellular Vesicles (EVs) and Exosomes (Exos)

EVs partake in neuroinflammation-promoting intercellular signaling. Exos are a class of EVs extruded by any cell type. Exos originate in multivesicular bodies, have sizes of 30–100 nm, and bear specific tetraspanin family markers on their membranes. Exos enclose and convey high numbers of functional proteins, lipids, and regulatory RNAs, which affect recipient cells’ metabolic activities, proliferation, or death. Hence, nerve cell-released Exos can act as “either friends or foes” to neurons depending upon their cargoes (e.g., growth factors or Aβs or p-Taues) [60,208] (v. Table 2). In a model of microglial BV-2 cells, pyroptosis induced by O2-glucose deprivation/reperfusion (OGD/R), human mesenchymal stem cells (MSC)-released Exos (huMSC-Exos) increased FOXO3a gene expression, thereby enhancing mitophagy while reducing the levels of NLRP3; cleaved caspase-1, IL-1β, IL-18; GsdmD-N fragments; and pyroptosis. Hence, huMSC-Exos might mitigate human neurons’ OGD/R-induced pyroptosis [209]. Consistently, bone marrow MSC-derived Exos (BMMSC-Exos) intravenously injected 2 h after middle cerebral artery occlusion (MCAO) decreased brain infarct volume, NLRP3 protein expression, and neuron pyroptosis. Moreover, BMMSC-Exos administration shifted the ischemia-induced microglial proinflammatory M1 phenotype to the homeostatic M2 [210].
Cui et al. [197] reported that Exos released from hypoxia-preconditioned MSCs (MSC-Exos) downregulated TNF-α and IL-1β, hindered NF-κB and STAT3 (signal transducer and activator of transcription 3) activation, and decreased Aβ peptides levels and senile Aβ plaques, while upregulating anti-inflammatory IL-4 and IL-10, and exo-miR-21, which improved memory and learning in APP/PS1 AD-model mice. In another study, Cui et al. [211] used the CNS-specific rabies viral glycoprotein (RVG) to target intravenously infused Exos released from MSCs (MSC-RVG-Exos) to the cerebral cortex and hippocampi of transgenic APP/PS1 AD-model mice. MSC-RVG-Exos downregulated IL-1β, TNF-α, and IL-6, while upregulating anti-inflammatory IL-10, IL-4, and IL-13.
In summary, the available evidence about EVs’ and Exos’ beneficial or harmful roles in NLRP3-mediated neuroinflammation is still scanty. A further limitation is that most studies focused on the RNAs conveyed by EVs or Exos. However, EVs or Exos also transport high numbers of different proteins that either promote or hinder neuroinflammation. In fact, Exos from Aβ25–35-exposed human cortical astrocytes conveyed significantly increased amounts of p-Taues [212], while Exos from human AD brains transported Aβ oligomers [213].

1.5. Other Brain NLRP3 Inflammasome Regulators

Under any situation, complex sets of endogenous factors control or restrain NLRP3 inflammasome assembly and/or function, trying to reestablish and/or upkeep tissue homeostasis. Zhang et al. [214] strengthened the relevance of the NLRP3 concept by proving that NLRP3 gene knockout or pharmacological blockage improved the course of various inflammatory diseases modeled in rodents. Hereafter we mention relevant NLRP3 regulators.
The zinc-finger protein A20, i.e., TNFAIP3 (TNF-α-induced protein 3), has two functions: it blocks apoptosis and crucially controls microglia function by inhibiting NF-κB activation in CNS physiological and pathological conditions. A20 knockout led to NLRP3 inflammasome’s hyperactivation, increasing mature IL-1β secretion and neuroinflammation intensity [215].
Additionally, CD40 (i.e., cluster of differentiation 40) protein, a member of the TNFR superfamily, negatively affected the ATP•TLR4-signaling-mediated NLRP3 inflammasome’s activation in microglia. Therefore, it regulated microglia’s inflammation-initiating Th17 response triggered by DAMP-induced brain injuries [216].
Mitsugumin-53 (i.e., TRIM-72 or tripartite motif 72) protein partook in damaged plasma membranes repair and inhibited the NLRP3/caspase1/IL-1β pathway and TNF-α expression, thus mitigating neuroinflammation [217]. Conversely, the TRIM-21 protein promoted microglia’s pro-inflammatory M1 phenotype polarization that TRIM-21’s knockout reversed [218].
Osteopontin is a highly phosphorylated ECM sialoprotein expressed during the subacute phase following cerebral infarction. It stimulated microglia’s chemotaxis while preventing NLRP3’s activation and its sequels [52].
Worth mentioning here is PKR (i.e., protein kinase RNA-activated), a multirole serine–threonine kinase controlling mRNA transcription/translation, protein synthesis, cell proliferation, apoptosis, and brain function, in addition to shielding cells from viral infections. A dysfunctional PKR partook in cancer and neuroinflammation [219]. Moreover, by using wild-type and PKR−/− mouse macrophages, Lu et al. [220] showed that PKR needed to physically interact with NLRP3, NLRC4, and AIM-2 inflammasomes to activate them. However, using LPS-treated PKR−/− bone marrow-derived macrophages isolated from different mouse strains, He et al. [221] reported that following stimuli activating NLRP3, NLRC4, and AIM2 inflammasomes’ PKR activity was critical for nitric oxide synthase-2 (NOS-2) induction, yet dispensable for pro-IL-1β and pro-IL-18 cleavage by caspase-1 [172]. Altogether the divergent results of Lu et al. [220] and Healy et al. [172] show that the animal species or strains investigated do significantly affect the kind of mechanisms activating or inactivating the NLRP3 and other inflammasomes. This adds a remarkable degree of complexity to the topic and stresses the importance of investigating corresponding mechanisms in human neural cells models.

1.6. Brain NLRP3 Inflammasome Inhibitors

Inhibiting the NLRP3 inflammasomes has been a tantalizing enterprise given its potential therapeutic applications in brain diseases. Table 3 lists the reported NLRP3 inhibitors, of which MCC950 is the most popular one in experimental works [222], although it failed in a clinical trial due to off-target toxic effects.

1.7. Brain NLRP3 Downregulation by Officinal Plant Agents/Herbal Extracts

Since time immemorial, plants were and still are the source of drugs helping human ailments. Although extracts of plant body portions are still in use in Traditional Chinese Medicine (TCM), the current more scientific attitude is to find the specific compound(s) of potential therapeutic use. Table 4 reports the most relevant agents and herbal extracts of interest regarding the brain NLRP3 inflammasome.
It is worth noting that save for ginsenoids, artemisinin, and artesunate, all the other hitherto-reported therapeutically promising plant agents/herbal extracts still need in-depth preclinical studies and well conducted clinical trials prior to becoming FDA-approved drugs. On the other hand, altogether the above-listed agents/extracts represent a treasure trove of future therapeutic assets.

2. NLRP3 Inflammasome in Brain Acute Injuries

Glial NLRP3’s role is controversial in HI/OGD (oxygen–glucose deprivation)-model animals. Denes et al. [336] reported that plasma IL-18 levels and brain infarction volume were alike in both wild-type and NLRP3-shRNA-silenced mice. Therefore, NLRP3’s downregulation was not as neuroprotective as expected because other inflammasomes took over and functioned in NLRP3’s stead. In fact, after shRNA-induced NLRP3 depletion, OGD significantly increased AIM2 inflammasome’s expression while NLRC4’s expression did not change in BV-2 microglial cells.
Conversely, Yang et al. [337] showed that in newborn mouse astrocytes HI and OGD activated TRPV1 (transient receptor potential vanilloid 1), a non-selective cation channel of the TRP family. Next, the TRPV1 signaling drove the JAK2-STAT3 pathway, which mediated NLRP3 inflammasome’s activation and increased IL-1β levels. Notably, in HI- and OGD-exposed TRPV1−/− mouse astrocytes, JAK2 and STAT3 activation and IL-1β upregulation were less intense. Interestingly, this study revealed different cell type-related timings of NLRP3 activation elicited by HI/OGD. In newborn mouse astrocytes of the hippocampus, striatum, and thalamic habenula, NLRP3’s activity increased by 3 h, while in microglia it was insignificant at 3 h but increased remarkably by 72 h. Then again, Schölwer et al. [338] showed that OGD completely inactivated phagocytic activity in wild-type BV-2 cells, while HI restored phagocytic activity in NLRP3-shRNA-depleted BV-2 cells. Therefore, the authors posited that NLRP3 plays a minor replaceable role in the OGD-elicited neuroinflammation, at least in microglia. Conversely, an anti-inflammatory pleiotropic cytokine, IL-10, hindered NLRP3 activation in microglia by increasing STAT-3’s function, which stifled the transcription/translation of pro-IL-1β and mature IL-1β production [339].
Relevant to this topic is IL-33, another IL-1 family member playing major pleiotropic roles in normal and pathological conditions [340]. In neonatal mouse astrocytes, IL-33 expression markedly increased by 24 h after a cerebral HI episode. Exogenously administered IL-33 did mitigate brain infarction volume by one week after the HI event. Astrocytes’ basal expression of ST2 (or suppressor of tumorigenesis 2), the IL-33 receptor, was intense and after HI exposure increased further. Conversely, a ST2 shortfall worsened the HI-elicited brain infarction. The IL-33•ST2 signaling-activated pathways mitigated astrocytes’ HI-elicited neuroinflammatory response and apoptosis. Moreover, in vitro IL-33-treated murine astrocytes released neurotrophic factors, which protected HI- and OGD-exposed neurons’ viability [341]. Besides, administering IL-33 plus MCC950 and antimalarial drugs improved the outcome in a model of murine cerebral malaria [342] in which the Plasmodium falciparum overgrew inside the cortical capillaries, diffusely obstructing blood flow.
Franke et al. [36] showed that following stroke’s onset, the early up-regulation of the NLRP3 inflammasome occurred in neurons, glia, and vascular endothelia, leading to blood–brain barrier (BBB) breakdown. Consistently, NLRP3 inhibition hindered endothelial pyroptosis induced by the thrombolytic agent rt-PA (or tissue plasminogen activator), thus preserving the BBB’s integrity [11]. Similarly, NLRP3-inhibitor MCC950 protected brain endothelial cells from rt-PA’s toxic effects in an in vitro HI-exposed BBB model [343]. Additionally, NLRP3′s knockout alleviated the NF-κB pathway-mediated brain damage in a middle cerebral artery occlusion (MCAO)-induced focal ischemia mouse model [344]. Moreover, lithium (Li+), the archetypal mood stabilizer, also impeded HI/R-induced NLRP3 inflammasome activation, and by stimulating STAT3’s function improved motor behavior, cognition, and depression [345].
Figure 1 sums up the main signaling pathways involving NLRP3 in acute brain injuries.
Finally, electroacupuncture (EA) exerted analgesic effects by suppressing NLRP3 inflammasome function in the spinal dorsal horn of mice [346]. Moreover, EA at the skull’s Shenting (DU24) and Baihui (DU20) acupoints attenuated cognitive impairment in rats with brain HI/R injury by regulating endogenous melatonin secretion through alkylamine N-acetyltransferase synthesis in the epiphysis. Next, melatonin acted neuroprotectively by blocking NLRP3 activation via upregulating mitophagy-associated proteins [347].
In conclusion, given the consistent risk that an acute brain injury triggers a chronic neurodegenerative disease entailing a lethal outcome, the therapeutic mitigation or better suppression of neuroinflammation within a brief time lag following the harmful event constitutes a quite valid target to be pursued.

3. NLRP3 Inflammasome in Chronic Neurodegenerative Disease

3.1. Alzheimer’s Disease (AD)

AD is the most prevalent human dementia. Under healthy conditions, the NLRP3 inflammasome is inactive in microglia and astrocytes. Halle et al. [348] first showed that Aβ fibrils—AD’s main drivers together with p-Taues and neuroinflammation—activate microglia’s NLRP3 inflammasome in APP/PS1 AD-model mice. After phagocytosis by primary mouse microglia, Aβ1–42 fibrils damaged the lysosomes, which released cathepsin B, activating the NLRP3 (previously named NALP3) inflammasome and IL-1β, TNF-α, and nitric oxide (NO) overproduction. In turn, the activated NLRP3 inflammasome intensified AD neuropathology in vivo well before Aβs senile plaques appeared [348,349,350]. Heneka et al. [349] also showed that NLRP3 inflammasome’s downregulation shifted microglia’s polarization toward the homeostatic M2 phenotype, concurrently depleting the brain’s Aβs load. Hence, they posited that NLRP3 inflammasome activation remarkably partook in the microglia-mediated persistent neuroinflammation observed in AD-model mice. Consequently, NLRP3’s inhibition would be a novel anti-AD therapeutic approach. Consistently, NLRP3-blocking dihydromyricetin [239] or MCC950 [222] promoted the brain’s Aβs clearance, increased hippocampal and cortical M2 microglia fractions, and improved memory and cognition in APP/PS1 mice.
Astrocytes are by far the most abundant cell type populating the brain. Hence, any astrocytes’ contributions to neuroinflammation are quite relevant to the progression/outcomes of neurodegenerative diseases. ASC is an adaptor protein forming stable NLRP3•ASC complexes acting as inflammasomal activation hubs. Studies using ASC+/− or ASC−/− 5xFAD newborn mice proved that Aβs do activate astrocytes’ inflammasome(s). In ASC+/− mice, NLRP3 inflammasome activity was downregulated; concurrently, an upregulated MIP-1α/CCL3 release increased Aβs phagocytosis by lipopolysaccharide (LPS)-primed primary newborn 5xFAD mouse astrocytes. Moreover, in 7–8-month-old ASC+/− 5xFAD mice, Aβs’ brain load downfall correlated with upregulated CCL3 gene expression and improved spatial reference memory [351,352]. Furthermore, ASC moieties released from pyroptotic neurons bound extracellular Aβs and cross-seeded Aβs’ increase, promoting NLRP3 inflammasome’s activation, neuronal pyroptosis, and neuroinflammation. In turn, these effects increased ASC’s available moieties, triggering a self-sustaining feedforward vicious loop while undermining microglial Aβs clearance [353].
Murphy et al. [354] showed that exposure to Aβs increased cytosolic cathepsin B’s protease activity, which drove NLRP3 inflammasome’s activation and IL-1β over release from wild-type rat primary glial cultures. Consistently, the endogenous protease inhibitor α1-antitrypsin (A1AT) reduced Aβ1–42-elicited NLRP3’s activation and its sequels in primary cortical astrocytes from BALB/c mice [128,222].
More recent investigations using rodent astrocytes confirmed that exposure to Aβ1–42 or LPS inhibited the autophagy/lysosome function while activating the NLRP3/ASC/caspase-1/IL-1β pathway. However, the administration of rapamycin or 17β-estradiol (E2) or progesterone rescued autophagic activity while curbing the Aβ1–42- and LPS-activated NLRP3/caspase-/IL-1β pathway in the astrocytes. By contrast, 3-methyladenine, a specific autophagy inhibitor, blocked progesterone’s neuroprotective effects and drove astrocytes’ NLRP3 inflammasome activation and neuroinflammation [355,356].
Here, a mention is in order about the inducible thioredoxin-interacting protein (TXNIP), which partakes in oxidative stress and regulates thioredoxin (TRX), another redox controller. Both the unfolded protein response (UPR) and ER stress also activate TXNIP. Concurrently, UPR activates the IRE-1α (or inositol requiring enzyme-1α) stress sensor pathway, which in turn further increases TXNIP’s amounts susceptible of activation [357]. Importantly, TXNIP’s function is essential for the increased expression and activation of NLRP3’s inflammatory cascade, both in the aging-associated chronic inflammaging, which goes along with senile cognitive decline, and in the hippocampal neurons and microglia of AD brains [66,67,358]. In rodent models of AD, Aβ1–42 drove NLRP3 activation and oxidative damage via the formation of TXNIP•Keap1 (Kelch-like ECH-associated protein-1)•NRF2 (nuclear factor erythroid 2-related factor 2) complexes. Exposure to HJ105 or HJ22, both piperine derivatives, or 9-(NXPZ-2) or maxacalcitol, an active vitamin D analogue, directly inhibited the formation of Keap1•NRF2 complexes, upregulated NRF2’s nuclear expression, hindered TXNIP-mediated NLRP3 inflammasome activation, and blocked Aβ1–42 and oxidative stress noxious effects [359,360,361,362].
Figure 2 sums up the main signaling pathways involving NLRP3 in AD.
Notably, ER stress concurs with the depletion of the anti-aging and cognition-enhancing Klotho, FOXO-1, and mTOR proteins. Moreover, proteins partaking in ER stress development—such as BiP (binding immunoglobulin protein), eIF-2α (eukaryotic initiation factor-2α), and CHOP (C/EBP homology protein)—showed heightened levels of expression in the hippocampi of AD brains. Therefore, altogether TXNIP could link the chronic increases in glucocorticoids elicited by a persistent ER stress with AD’s enduring NLRP3 activation and neuroinflammation [67,363].
A newly identified gene associated with the risk of AD is TREML2 (triggering receptor expressed on myeloid cell-like 2), a protein expressed by microglia [364,365]. TREML2 protein expression levels rise along with AD progression in vivo [366] and after LPS stimulation in primary microglia in vitro, both proving TREML2 involvement in microglia-induced neuroinflammation [367]. Then again, Wang et al. [368] showed that LPS stimulation or lentivirus-mediated TREML2 overexpression remarkably upregulated NLRP3 inflammasome activation; IL-1β, IL-6, and TNF-α secretion; and proinflammatory M1-type polarization in microglia of APP/PS1 AD-model mice. Therefore, TREML2 inhibition would be a novel anti-AD therapeutic approach.
Two studies showed that caspase-1-mediated overproduction of IL-1β occurred in brain samples from mild cognitive impairment (MCI) and fully symptomatic AD patients. Hence, in both groups, microglial NLRP3 inflammasome activation advanced AD’s persistent neuroinflammation [140,348]. Sokolowska et al. [140] also showed that phagocytosed Aβ1–42 fibrils damaged human macrophages’ lysosomes, which released cathepsin B into the cytosol, triggering the NLRP3•ASC•caspase-1 inflammasome’s oligomerization and activation. Moreover, studies conducted on brain tissue samples from AD patients that had died because of intercurrent systemic infections and APP/PS1 AD-model mice revealed that any added proinflammatory insults intensified NLRP3 inflammasome’s assembly/activation and IL-1β, IL-6, and various chemokines release from microglia, astrocytes, and neurons while increasing the brain’s Aβs and p-Taues load. Hence, any concurring etiologic factor could worsen neuroinflammation and hasten AD progression in humans [71,369,370].
Saresella et al. [371] reported the occurrence of a significantly upregulated expression of mRNAs encoding for NLRP1; NLRP3; ASC/PYCARD; caspase-1, -5, and -8; pro-IL-1β; and pro-IL-18 in monocytes isolated from MCI or late-stage AD patients. However, both NLRP1 and NLRP3 inflammasomes functioned only in late-stage AD monocytes. Conversely, ASC/PYCARD and caspase-1 expression was normal in early MCI monocytes in which assembled/functional inflammasomes were missing. Hence, concurrently activated NLRP1 and NLRP3 inflammasomes aggravated neuroinflammation only in late AD.
Interestingly, in subjects with autistic spectrum disorders (ASD), Saresella et al. [131] found that both AIM2 and NLRP3 inflammasomes were active, overproducing IL-1β and IL-18. Simultaneously, there occurred an upregulation of the innate immunity suppressor IL-37, a decline of anti-inflammatory IL-33, and a rise in IFABP (intestinal fatty acid-binding protein—an altered gut permeability index). Therefore, multiple inflammasomes are active in both AD and ASD.
Immunohistochemical studies conducted on samples of temporal cerebral cortex of AD brains showed that the increased expression of NLRP3 inflammasome’s constituents, including pro-caspase-1, and of IL-1β and IL-18, co-localized with glia maturation factor (GMF), APOE-ε4, sequestosome 1 (SQSTM1)/p62, LC3-positive autophagic vesicles, and LAMP1, a lysosomal marker. Notably, clusters of GMF overexpressing reactive astrocytes surrounded the amyloid senile plaques. GMF is a highly conserved proinflammatory protein that activates glial cells advancing human neurodegenerative processes. Conversely, in AD-model animals, GMF suppression mitigated the neurodegeneration. Altogether, these results showed that in humans, GMF could intensify NLRP3-driven neuroinflammation while concurrently hampering the autophagosomal pathway clearing Aβs aggregates [349]. Of note, Ahmed et al. [372] and Ramaswamy et al. [352] posited that GMF may advance neuroinflammation in all neurodegenerative diseases.
By sharp contrast, the results of another human postmortem study negated NLRP3 inflammasome function in the brains of advanced AD cases in which astrocyte activation was instead prominent [132].
In addition to Aβs and neuroinflammation, p-Taues are among AD’s main drivers. Stancu et al. [373] and Ising et al. [71] proved that a causal link existed between p-Taues and inflammasomes’ activation. They showed that following microglial endocytosis and lysosomal sorting, prion-like Tau seeds activated NLRP3 inflammasome signaling in the THY-Tau22 transgenic mouse line, a tauopathy-model animal. Moreover, the chronic intraventricular administration of NLRP3 inhibitor MCC950 significantly thwarted the neuropathology driven by the exogenous p-Tau seeds. Concurrently, NLRP3 suppression decreased the p-Taues levels and hindered their aggregation into neurofibrillary tangles by restraining Tau kinases’ activity while increasing that of p-Tau phosphatases [71]. Then again, Jiang et al. [60] showed that p-Tau paired-helical filaments and p-Taues from human tauopathy brains primed and activated IL-1β production via MyD88 and NLRP3•ASC•caspase-1 pathways in primary human microglia. The authors also showed that p-Taues accumulation concurred with elevated ASC and IL-1β levels in postmortem brains of tauopathies patients.
Autophagy is a conserved process by which lysosomes remove dysfunctional cellular components and relevantly regulate NLRP3’s role in inflammatory CNS diseases [10,374]. A reduced biogenesis and function of lysosomes/autophagosomes promotes the NLRP3’s inflammasome activation driving the neuroinflammatory response in AD-model animals and cultured neural cells. In keeping with this, Zhou et al. [375] showed that overexpressing the transcription factor EB (TFEB), the primary regulator of lysosomal biogenesis, both improved the autophagosomes/lysosomes function and mitigated the neuroinflammation in AD-model cells.
Summing up, NLRP3 inflammasome targeting might hinder AD’s etiopathogenetic tripod, i.e., Aβs, p-Taues, and neuroinflammation, and beneficially affect tauopathies too. This is indeed a sensible proposal, but hitherto its real effectiveness in stopping human AD’s progression is unproven. Moreover, it does not consider inflammasomes’ plurality, potential functional interchangeability, and their different expression levels in the distinct neural cell types.

3.2. Parkinson’s Disease (PD)

PD is the second-most-common age-related human neurodegenerative disorder. The progressive spread of PD neuropathology causes motor disturbances and neuropsychiatric disorders (e.g., depression). PD’s hallmarks are inclusions rich in misfolded α-synuclein (α-Syn) protein localized at the presynaptic terminals of melanin-rich dopaminergic neurons within the mesencephalic substantia nigra and subcortical corpus striatum. Zhang et al. [376] found the overexpression of IL-1β and IL-18 in cerebrospinal fluid samples from PD patients. Consistently, α-Syn mediated NLRP3 inflammasome activation in cultured human microglia [64]. In PD-model animals, β-hydroxybutyrate, a ketone body, did not inhibit NLRP3 [377] while blocking it in AD [378]. Therefore, α-Syn aggregates trigger chronic neuroinflammation sustained by mitochondrial dysfunction causing ROS overproduction and by unrestrained microglia activation advancing dopaminergic neurons’ pyroptosis [379,380,381].
Figure 3 sums up the main signaling pathways involving NLRP3 in PD.
Moreover, Scheiblich et al. [382] reported that the signaling triggered by the binding of α-Syn monomers or, to a lesser extent, α-Syn oligomers to TLR-2 and TLR-5 receptors activated the NLRP3 inflammasome in microglia with no priming needed. Using immunohistochemical and genetic approaches, von Herrmann et al. [383] supplied evidence that dopaminergic neurons are sites of NLRP3 activity in PD. Moreover, increases in NLRP3 inflammasome and NLRP3-dependent pro-inflammatory cytokines were detectable in the peripheral plasma of PD patients, proving NLRP3 inflammasome involvement in PD’s pathogenesis [196,384]. The latter authors also showed that miR-7 inhibited NLRP3 gene expression in microglia, thereby reducing microglia activation, neuroinflammation, and nigrostriatal dopaminergic neuron pyroptosis. A patient-based study characterized NLRP3 in the first stages of midbrain nigral neurodegeneration and in the biofluids drawn from PD patients, suggesting that NLRP3 may be both a key inflammation mediator in the degenerating midbrain and a tractable therapeutic target [385]. Moreover, Wang et al. [386] showed that NLRP3 activation and IL-1β and IL-18 maturation occurred in the 6-OHDA (6-hydroxydopamine) neurotoxin-induced PD-model rat. The purinergic P2X4-R siRNA-knockdown or block by the specific antagonist 5-BDBD (5-(3-bromophenyl)-1,3-dihydro-2H-benzofuro[3,2-e]-1,4-diazepin-2-one) counteracted NLRP3’s effects, alleviated neuroinflammation, and reduced dopaminergic neuron pyroptosis. Therefore, the authors posited that the ATP•P2X4-R signaling drives NLRP3 inflammasome’s activation, which next regulates glial cell activation, nigrostriatal dopaminergic neurodegeneration, and dopamine levels (Figure 3; see also more details and the literature in Box 2) However, here one should be wary of extrapolating these data to PD patients. In PD-model rat brains, NLRP3 inflammasome’s activation is not in fact equivalent to that proper of human PD brains. The present understanding of any beneficial effects of antagonizing ATP•P2X4-R’s signaling is too limited. Therefore, we need more studies to assess the pathophysiological relevance of nigrostriatal ATP•P2X4R signaling in humans.
Consistently, inhibiting NLRP3 function with MCC950 evoked substantial neuroprotection in the 6-OHDA PD-model rats [387] and in MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine)-induced PD-model mice [384]. Moreover, NLRP3 inflammasome’s activation in microglia promoted the extracellular release of α-Syn-conveying Exos, which could advance α-Syn spreading in PD brains [388]. Interestingly, copper (Cu2+) accumulation also advanced PD’s pathogenic mechanisms by inducing ROS-mediated oxidative stress, activating the NF-κB-p65 pathway in BV2 microglial cells [49]. A persistent intracellular Cu2+ buildup upregulated the NLRP3 pathway-related proteins, advancing proinflammatory cytokine secretion and a disordered mitochondrial autophagy (or mitophagy), altogether resulting in dopaminergic neuron pyroptosis. Of note, Cu2+ may drive AD neuropathology as well [48].
Finally, despite extensive investigations into the NLRP3 inflammasome-activating mechanisms in the diverse inflammatory brain diseases, their regulatory networks are still unclear in microglia and other neural cell types. Chen et al. [389] showed that NLRP3 is a substrate of chaperone-mediated autophagy (CMA). The p38/TFEB (transcription factor EB) axis regulated NLRP3 inflammasome degradation via CMA, inhibiting the overproduction of proinflammatory cytokines in microglial cells. Furthermore, both p38 and NLRP3 inhibitors could mitigate α-Syn aggregate-induced microglia activation and nigrostriatal dopaminergic neuron pyroptosis. Moreover, Panicker et al. [390] showed that the functional loss of Parkin, an E3 ubiquitin ligase, resulted in the priming and spontaneous activation of the NLRP3 inflammasome in mouse and human dopaminergic neurons, leading to their pyroptosis.
From a clinical standpoint, human PD is quite complex. Therefore, one may conclude that the roles of NLRP3 and other-than-NLRP3 inflammasomes in human PD require further investigations to be fully clarified and integrated to lead to effective therapeutic interventions.

3.3. Multiple Sclerosis (MS) and Experimental Autoimmune (or Allergic) Encephalomyelitis (EAE)

MS is a chronic autoimmune disease of unclear etiology affecting both the brain and spinal cord whose hallmarks include focal (plaque) demyelination and chronic neuroinflammation/neurodegeneration. One accredited theory posits that patients’ T cells attack myelin sheath antigens, causing MS. The suggested relationship between MS and the NLRP3 inflammasome has linked autoimmunity with innate immunity and neuroinflammation [391,392,393,394,395]. Moreover, as gain-of-function genetic variants of the NLRP3 (e.g., Q705K) and NLRC4 inflammasomes associate with a more severe MS course, a constitutive NLRP3 inflammasome activation could be a risk factor for clinical MS presentation [396]. Moreover, Vidmar et al. [397] highlighted as pathogenetically important for MS patients the increased burden of rare variants in (i) NLRP1 and NLRP3 genes; (ii) genes partaking in inflammasome downregulation via autophagy and IFN-β; and (iii) genes involved in responses to type-1 IFNs (e.g., PTPRC, TYK2) and to DNA virus infections (e.g., DHX58, POLR3A, IFIH1).
Keane et al. [398] and Voet et al. [215] showed that following NLRP3 inflammasome activation, there occurred an increased IL-1β gene expression within MS demyelination plaques coupled with elevated levels of ASC, caspase-1, and IL-18 in the brains and cerebrospinal fluids of MS patients. Moreover, NLRP3 inflammasome pathway-related components were overexpressed in the blood monocytes isolated from the minor fraction of patients suffering from primary progressive (i.e., with no alternation of pauses and relapses) MS (PPMS), so entailing increased IL-1β production [393,394,399]. These results showed IL-1β as a prognostic factor in PPMS patients and the NLRP3 inflammasome as a prospective therapeutic target. Thus, a specific NLRP3 inhibitor may improve MS histopathology and reduce myelin sheath damage.
According to Farooqi et al. [400], EAE is a proper mouse model for pathogenetic and pharmacotherapeutic studies into human MS molecular mechanisms. In EAE-model mice, NLRP3 inflammasome’s activation critically induced T-helper cell migration into the CNS. Next, the activated NLRP3 inflammasome of primed T cells (and microglia) drove the release of proinflammatory cytokines, thus partaking in MS pathogenesis [394,401]. In EAE-model mice the NLRP3 inhibitor MCC950 prevented the conversion of CNS astrocytes to the A1 neurotoxic reactive phenotype otherwise induced via the NF-κB pathway-mediated IL-18 production. Consistently, after the systemic delivery of NLRP3 inhibitor MCC950 axonal injury was mitigated within lysolecithin-induced demyelinated lesions in mice [402,403]. MCC950 also hindered complement C3 protein release from the astrocytes, which would have otherwise impaired hippocampal neuron viability [404]. IFN-β administration did improve this NLRP3-dependent EAE form. Conversely, when ad hoc experimental regimens brought about a NLRP3-independent, more aggressive EAE, the IFN-β treatment was ineffective. A similar NLRP3-independent mechanism might be at work in human MS cases not profiting from IFN-β therapy [405].
In conclusion, there is an intensely felt need to expand the study of NLRP3 and other-than-NLRP3 inflammasomes’ role(s) in MS, using human neural cell-based experimental models to achieve a more detailed molecular picture and identify disease-modifying therapeutic targets.

3.4. Amyotrophic Lateral Sclerosis (ALS)

ALS is a devastatingly progressive multifactorial disorder characterized by the primary degeneration of the cerebral motor cortex, brain stem, and spinal cord motoneurons leading to skeletal muscle atrophy and paralysis. ALS patients may also develop cognitive and behavioral changes due to neurodegeneration-affected subcortical areas, e.g., diencephalon’s dorsal thalamus. Typically, 90% of cases occur sporadically, and their etiological factors are poorly defined (smoking, violent sports, military service, exposure to insecticides and pesticides). About 10% of ALS cases are familiar due to heritable mutated genes. SOD1 (superoxide dismutase 1) gene mutations occur in 20% of familiar cases [406]. The current belief is that SOD1 mutations only trigger ALS onset within motoneurons but elicit only delayed and minor harm [407]. However, in astrocytes and/or microglia, SOD1 mutations advance ALS progression [408]. TDP-43 (transactive response DNA binding 43 kDa) protein could be another ALS etiological agent as it accumulates in both sporadic and familial cases [409]. TDP-43 forms toxic ubiquitinated aggregates in the cytoplasm of neural cells of both ALS and frontotemporal lobar degeneration (FTLD) patients [410,411]. Neurons and astrocytes can secrete mutated or oxidized SOD1 and TDP-43 as misfolded proteins, which activate microglia by interacting with CD14, TLR-2, TLR-4, and scavenger receptors [412,413]. Thus, exogenous whole or fragmented, wild-type or mutated TDP-43 bound microglia’s CD14 cell surface receptor activating AP1 and NF-κB pathways and upregulating NOX2 (SOD-generating NADPH oxidase 2), TNF-α, NLRP3•ASC•caspase-1, and IL-1β release. Importantly, TDP-43 was toxic to motoneurons only in the presence of microglia presence [414]. Using in situ hybridization and immunocytochemistry, Banerjee et al. [415] showed that an upregulated NLRP3 inflammasome occurred in neurons and glia of cognitively impaired ALS patients. Conversely, no differences were detectable between cognitively resilient ALS and healthy subjects.
Figure 4 sums up the main signaling pathways involving NLRP3 in ALS.
Johann et al. [127] showed that an activated NLRP3 inflammasome concurred with elevated levels of caspase-1, IL-1β, and IL-18, particularly in the spinal cord astrocytes of the SOD1G93A ALS-model mice and in the serum and spinal cord tissue of sporadic ALS patients—altogether findings confirming NLRP3 inflammasome’s involvement in ALS. Moreover, Kadhim et al. [416] found that IL-18 was upregulated in the cerebral tissue of sporadic ALS patients vs. age-matched controls. Furthermore, Gugliandolo et al. [417] strengthened the concept that neuroinflammation plays a crucial role in ALS by confirming NLRP3 inflammasome activation and its sequels in SOD1G93A ALS-model rats. Immunofluorescent studies conducted on symptomatic SOD1G93A ALS-model mice revealed that NLRP3 and ASC expression intensity increased along with ALS progression, proving NLRP3’s involvement in neuron death [418]. Moreover, Michaelson et al. [419] suggested a novel ALS pathogenetic mechanism mediated by the amino acid β-N-methylamino-l-alanine (BMAA), a Cyanobacteria product. BMAA is not a protein constituent, but a powerful neurotoxin inducing protein misfolding, NLRP3 inflammasome activation, and proinflammatory cytokine overexpression in spinal motoneurons.
In their work, Van Schoor et al. [420] observed increases in the NLRP3 inflammasome, GSDMD-N fragments, and IL-18 in the motor cortex and spinal cord microglia of human ALS patients, which suggested that an activated NLRP3 inflammasome had triggered the cells’ pyroptosis. As compared to controls, in human ALS samples, a reduced array of neurons matched with an increased throng of cleaved-GSDMD-positive microglial cells in the underlying white matter of the premotor cortex. No alike findings were obtained in the human spinal cord. Similar findings were made in the cortex of TDP-43A315T transgenic mice in model ALS and FTLD [421]. In addition, these results stressed the relevance of ROS and ATP generation, both potential therapeutic targets, for microglial NLRP3 inflammasome activation and neuronal pyroptosis, which was confirmed in SOD1G93A-induced ALS-model mice. Importantly, both wild-type and mutant TDP-43 proteins activated the overexpressed NLRP3 and its downstream effects in the microglia of SOD1G93A mice. This proved that NLRP3 is the crucial microglial inflammasome mediating SOD1G93A-induced pyroptosis [65].
Lacking a suitable human microglia model, Quek et al. [422] characterized peripheral blood monocyte-derived microglia-like cells (ALS-MDMi) isolated from ALS patients at various stages. Importantly, ALS-MDMi recapitulated ALS neuropathology hallmarks, i.e., abnormal phosphorylated and non-phosphorylated TDP-43 cytoplasmic accumulation and phagocytosis impairment that paralleled ALS progression; altered neuroinflammatory cytology; DNA damage; NLRP3 inflammasome’s activation; and microglia pyroptosis.
It is seemly to consider the studies about NLRP3 and other-than-NLRP3 inflammasomes in human ALS are still in a preliminary phase even in the light of the groundbreaking results reported by Van Schoor et al. [420]. The latter should encourage scientists to delve deeper into the pathogenetic mechanisms of this devastating disease to find novel effective therapeutic approaches.

3.5. Huntington’s Disease (HD)

HD is a rare autosomal dominant neurodegenerative disease caused by the unstable CAG repeat expansion in the Huntington (HTT/IT15) gene and presenting with motor, cognitive, and psychiatric symptoms [423] When the HTT/IT15 gene holds 39 to 180 CAG repeats, the translated polyglutamine-containing mutant HTT protein (mHTT) complexes with and disrupts the normal function of several transcription factors, thereby altering the activities of neurons, astrocytes, and microglia. HD’s harming mechanisms include mitochondrial dysfunction, excitotoxicity, CREB and BDNF downregulation, and microglia activation, altogether advancing neuronal death by apoptosis, necroptosis, ferroptosis, and NLRP3-linked pyroptosis [424,425].
Various HD-model animals were set up to clarify its molecular mechanisms and to try novel therapeutics for it. The transgenic R6/2 (B6CBA-Tg[HDexon1]62Gpb/1J) mouse line expressing the human HTT gene exon 1 carrying 120 ± 5 CAG repeats is the most popular HD animal model [426]. An upregulated NLRP3 inflammasome and caspase-1 expression already occurred in 13-week-old R6/2 HD-model mice, particularly in striatal parvalbumin interneurons and spiny GABAergic neurons, which preferentially undergo pyroptosis in HD [427]. Poly(ADP-ribose) polymerase-1 (PARP-1) is a nuclear enzyme whose activity is crucial for DNA repair in humans. Olaparib, a PARP-1 inhibitor presently sold as an anti-tumor drug, could also regulate NLRP3 inflammasome activation in the R6/2 HD-model mice. When given from the pre-symptomatic stage onwards, Olaparib mitigated neuronal pyroptosis, neurological symptoms, and neurobehavioral tests results, lengthening the survival of HD-model mice. Therefore, Olaparib could help human HD too [428]. Moreover, Chen et al. [429] showed that NLRP3 inhibitor MCC950 given to R6/2 HD-model mice suppressed IL-1β and ROS overproduction, mitigating neuroinflammation, motor dysfunction, and neuronal pyroptosis, while upregulating PSD-95 and NeuN proteins, and lengthening animals’ lifespans. Therefore, inhibition of NLRP3’s signaling, and its downstream effects would be therapeutically helpful in HD.
Interestingly, a role in HD etiopathogenesis may be played by galectins (i.e., “S-type lectins”)—soluble proteins specifically binding β-galactoside carbohydrates and playing multiple roles in autophagy, immune responses, and inflammation. Siew et al. [430] reported that galectin-3 (Gal-3) plasma levels increased well over healthy controls in HD patients and HD-model mice. In HD-mice, microglia Gal-3 levels increased prior to motor symptom presentation and stayed high while HD progressed. Gal-3 co-localized with microglial lysosomes, blocked the autophagic elimination of damaged endolysosomes, and partook in neuroinflammation via the NF-κB/NLRP3 axis. Gal-3 knockout improved HD-related neuropathology and survival in HD-model mice, showing Gal-3 as a potential therapeutic target. Conversely, Gal-1 and Gal-8 hindered neuroinflammation, promoting neuroprotective effects [431].
HD’s rare occurrence is an adjunct hurdle to studies about the roles played in it by NLRP3 and other inflammasomes. However, this circumstance should not discourage attempts to increase our insights in this ailment, both in patients and animal models.

4. Brain NLRP3 and Neurotropic Viruses Infections

Both DNA and RNA neurotropic viruses activate the brain’s NLRP3 inflammasome, causing neuroinflammation and sometimes triggering chronic neurodegenerative diseases [75]. Here, we review a few neurotropic viruses playing NLRP3-linked roles in human neuropathology.

4.1. Zika Virus (ZIKV) Encephalitis

The Zika Virus (ZIKV) is a single-stranded positive-sense RNA arbovirus of the Flaviviridae family (Flavivirus genus that also includes Dengue, West Nile, Yellow Fever, and Japanese Encephalitis viruses). ZIKV associates with congenital microcephaly in newborns and Guillain–Barré syndrome, myelopathy, and encephalitis in adults. Tricarico et al. [432] showed that ZIKV infected the U87-MG glioma cell line causing NLRP3 inflammasome activation and IL-1β oversecretion. Consistently, He et al. [82] made the same observations in the brains and sera of ZIKV-infected mice. ZIKV’s NS5 protein drove ROS overproduction and NLRP3 inflammasome assembly, both needed for its activation. Conversely, in vitro and in vivo NLRP3 deficiency upregulated type-I IFN and strengthened the host’s resistance to ZIKV, confirming NLRP3’s role in ZIKV infection [433,434].

4.2. West Nile Virus (WNV) Encephalitis

Another Flavivirus, the West Nile Virus (WNV), causes an encephalitis entailing neurons’ death and elevated IL-1β plasma levels. In a mouse model, WNV infection briskly induced IL-1β synthesis in cortical neurons. However, by cooperating with type-I IFN, the intensified IL-1β•IL-1β-R (receptor) signaling suppressed neuronal WNV replication, reducing the WNV brain load. Therefore, the NLRP3/IL-1β•IL-1β-R pathway regulated neuronal WNV infection and revealed a novel IL-1β antiviral action [435].

4.3. Japanese Encephalitis Virus (JEV)

By breaking the BBB, the Japanese Encephalitis Virus (JEV) enters the CNS where it induces a diffuse neuroinflammation. Thus, JEV infection activated (i) a ROS-dependent Src/Ras/Raf/ERK/NF-κB signaling axis in neurons/glia co-cultures [81]; (ii) a ROS/Src/PDGFR/PI3K/Akt/MAPK/AP-1 axis [436] and a PAK4/MAPK/NF-κB/AP-1 axis [437] in rat brain astrocytes; and (iii) via TLR-3 and RIG-I the ERK/MAPKp38/AP-1/NF-κB axis, ROS overproduction, and K+ efflux in cultured mouse microglia. These effects both triggered NLRP3 inflammasome signaling and polarized microglia toward the proinflammatory/neurotoxic M1 phenotype. In all instances, JEV advanced cytokine overproduction and neural cell pyroptosis [438].

4.4. Human Immunodeficiency Virus-1 (HIV-1) Encephalitis

The immunosuppressive Lentiviruses efficiently infect macrophages and lymphoid cells. Human Immunodeficiency Virus-1 (HIV-1) belongs to the Retroviridae family (Lentivirus genus). Burdo et al. [439] showed that during the primary infection, HIV-1 productively infects brain macrophages and microglia. Studies using primary human microglia showed that IL-1β was released after HIV-1 infection. Walsh et al. [440] proved that HIV-1 infection induced an NLRP3 inflammasome-dependent ASC translocation, caspase-1 activation, and mature IL-1β release from cultured microglia. The authors highlighted the need to analyze the inflammasome inhibitors’ effectiveness as novel therapeutics for HIV-1/AIDS.

4.5. Viroporin Proteins

Various RNA viruses, including Coronaviridae, express the viral-replication-indispensable small viroporin proteins. Being liposoluble, viroporins assemble hydrophilic transmembrane pores, allowing ions and/or small solutes to bidirectionally migrate along their electrochemical gradients. Viroporin activity could act as the “second signal” by increasing [Ca2+]i  or lowering the cytosolic pH due to H+-releasing ion channel activity in the lysosomal acidic compartment [441].

4.6. Encephalomyocarditis Virus (EMCV)

The Encephalomyocarditis Virus (EMCV) of the Cardiovirus genus (Picornaviridae family) is a non-enveloped, positive single-stranded RNA virus. Via an unclear sensing mechanism, the NLRP3 inflammasome detects EMCVs [442,443]. In this regard, Ito et al. [89] reported that by releasing Ca2+ from intracellular stores into the cytosol, ECMV’s viroporin ORF2b (or open reading frame 2b) triggered NLRP3 inflammasome activation.

4.7. SARS-CoV-2 Encephalitis

SARS-CoV-2 belongs to the β-Coronavirus genus (Coronaviridae family, also including 2003 SARS-CoV and 2012 MERS (Middle East Respiratory Syndrome)-CoV). SARS-CoV-2 is an enveloped single-stranded positive-sense RNA virus causing the COVID-19 (Coronavirus Disease 2019). The virus infects a wide spectrum of cell types. In the presence of Ca2+, SARS-CoV-2’s spike S1 glycoprotein binds ACE2 (angiotensin converting enzyme 2) and CD147 (cluster of differentiation 147) proteins, promoting virus endocytosis. Moreover, SARS-CoV-2’s envelope (E) protein binds TLR-2, which also helps promote AD and PD [444]. Earlier epidemics proved Coronaviruses’ neuroinvasive capability in humans [445,446]. SARS-CoV-2 infects neurons, astrocytes, microglia, and the BBB’s endothelial cells [447,448]. Notably, microglia and astrocytes are major sources of proinflammatory cytokines. Moreover, Sepehrinezhad et al. [449] found SARS-CoV-2 virions in the cerebrospinal fluid of COVID-19 patients presenting severe neurological symptoms previously affected or unaffected by neuropathologies and in “long COVID” patients [450]. However, in the healthy CNS, ACE2 expression is weak, prevailing in the brainstem’s respiratory centers—which explains the high prevalence of respiratory distress in COVID-19 patients [451]. However, uninfected AD patients showed upregulated ACE2 expression in the temporal and occipital neocortex and hippocampal CA1 subfield archicortex [452]. This ACE2 overexpression could advance SARS-CoV-2 infection in the same AD-hit inflamed areas, thus contributing to the high COVID19 mortality rates in aged AD patients [451].
Hitherto, SARS-CoV-2’s priming triggers are uncertain. Theobald et al. [453] showed that S1 spike glycoprotein initiated NLRP3 inflammasome activation. Other SARS-CoV-2 proteins—i.e., S, N, E, and the pore-forming viroporins ORF3a and ORF8—are also NLRP3 activators by causing K+ efflux and mitochondrial ROS over-release [83,84,85,86,87,88]. Moreover, Xu et al. [454] proved that viroporin ORF3a primed and activated the NLRP3 inflammasome through both ASC-dependent (canonical) and ASC-independent (noncanonical) pathways.
Notably, COVID-19 infection triggers a severe innate immune response producing elevated levels of multiple cytokines (“cytokines storm”) and inflammatory mediators (e.g., IL-1β, IL-2, IL-2-R, IL-4, IL-10, IL-18, IFN-γ, C-reactive protein, GCSF (granulocyte colony-stimulating factor), IP10, MCP-1, MIP-1α, and TNF-α). BV2 microglial cells exposed to SARS-CoV-2’s S1 spike glycoprotein expressed elevated levels of IL-1β, TNF-α, IL-6, NO, NLRP3, NF-κB signaling, and caspase-1 activity [88,455]. These cytokines cross the BBB inducing leukocyte infiltration, mitochondrial dysfunction, neuroinflammation, and neurons’ pyroptosis [442]. Interestingly, a mix of melatonin, vitamin C, and Zn2+ inhibited SARS-CoV-2-driven inflammasome activation, hindering the cytokine storm in animals [456].
Additionally, Ding et al. [457] proved that hypercapnia enhanced NLRP3 inflammasome activation and IL-1β expression only in hypoxic BV-2 microglia cells. Therefore, the hypercapnia resulting from lung-protective ventilatory strategies used in acute respiratory distress syndrome (ARDS) patients may lead to neuroinflammation and cognitive impairment via a microglial NLRP3/IL-1β-dependent mechanism.
Based upon the above findings, Heneka et al. [458] posited that NLRP3 inflammasome activation during COVID-19 heightens the risk for the later development of chronic neurodegenerative diseases. Independent clinical and epidemiological investigations indicated that SARS-CoV-2 infection and the ensuing “long COVID” tightly relate to the onset of AD, PD, prion disease (PrD), and other ailments, particularly in patients in advanced age or suffering from intercurrent illnesses (CVD, T2DM, hypertension, other neurological disorders) or severe/fatal COVID-19 [459,460,461]. Even more alarming, the receptor-binding domain of SARS-CoV-2’s S1 spike glycoprotein presents prion-like sequences. The latter diverge among viral variants, show a different affinity for ACE2, and promote immune-evasion, protein clustering, and protein aggregates’ “seeding”. The upshots would include prion-like proteins spreading, progressive dementia, or fast-evolving CJD [462,463,464].
Obviously, here we have considered only some of the known neurotropic viruses. The field of human brain-infecting viruses is more variegated and might also further expand in the future. Our knowledge about viral neuropathology is, we must admit, limited, particularly because viruses can target all stages of human life, from the uterus onward, with different age-related upshots. There is also a field that for the sake of brevity we omitted considering, i.e., the interactive relations between oncogenic viruses and inflammasomes, which deserves attention because of its potentially significant reflections on therapeutic outcomes.

5. Comments and Future Perspectives

An old dictum states that every disease starts with an inflammation. The prevalence of neuroinflammatory disease has been epidemically rising because of a lengthened lifespan and of little-appreciated toxic, environmental, and lifestyle-linked factors. To worsen this bleak situation, acute brain illnesses (e.g., stroke, hemorrhage, infection) too can trigger chronic neuroinflammation/neurodegeneration in a significant fraction of patients [465]. A steadily growing literature attests that NLRP3 inflammasome activation in CNS microglia and circulating monocytes plays a pivotal role in promoting the neuroinflammation driven by a host of etiologic factors (q.v. Table 1), potentially advancing the progression of neurodegenerative diseases [27,466,467]. Conversely, NLRP3’s roles in the other neural cell types (i.e., neurons, astrocytes, and oligodendrocytes) [3,468,469,470] and in CNS pericytes and endothelial cells [126,471] have received less attention, probably because such cells preferentially express other types of inflammasomes. In fact, NLRP3 activity in such cells is modest and/or is the object of controversy, particularly in astrocytes, although NLRP3’s inhibition still gives some therapeutic advantage. Moreover, these same neural cell types more intensely express various other-than-NLRP3 inflammasomes. The latter can also exert significant neuroinflammation-sustaining effects, as specific NLRP3 inhibitors do not hinder other-than-NLRP3 inflammasomes’ activities [24]. We previously reviewed the known roles of various other-than-NLRP3 inflammasomes in human brain disease [24]. That work inspired us to delve deeply also into the role(s) of the brain’s NLRP3 inflammasome. Indeed, the NLRP3-related extensive research works herein reviewed shows the high complexity of both the regulatory mechanisms involved and of the physiological, pathological, and ethnic/pharmacological factors that promote or hinder its activation. Particularly the abundance of blocking or preventative factors, many of them identified over millennia by TCM, bodes well for future therapeutic modulations of NLRP3 activity in various pathological settings. Various reports showed that particularly inhibiting microglial NLRP3 function exerted beneficial effects in rodent experimental models of human neurodegenerative illnesses. These favorable outcomes inspired and still inspire the opinion that therapeutically targeting the NLRP3 inflammasome will mitigate or stop both acute and progressive human neuroinflammatory diseases [472,473,474]. As just mentioned, despite or thanks to the intricacies of NLRP3 inflammasome’s activating mechanisms, there are plenty of agents modulating its activity (Table 2, Table 3 and Table 4). At present, many small molecules are undergoing pharmaceutical research/development as novel candidate drugs targeting the NLRP3 inflammasome in various diseases [274]. At least five companies have started ad hoc clinical trials, of which Inflazome and NodThera have reported Phase I positive results of their brain-penetrating NLRP3 inflammasome inhibitors (Inzomelid [251] and NT-0796 [274], respectively), expecting to use them to treat central and peripheral nervous inflammatory diseases. These discoveries have even raised the possibility of a common cure for all or at least some human brain diseases. Moreover, Lupfer and Kanneganti [21] reported the existence of inflammasomes, such as NLRC3, NLRP6, NLRP12, and NLRX1, which hinder NF-κB pathway activation, thereby mitigating or switching off the incumbent or ongoing neuroinflammation. Such “anti-inflammasomes” deserve more consideration because in a hopefully not too far future, their pharmacological activation by proper means (yet to be established) could be a valuable therapeutic asset that will switch off neuroinflammation through physiological mechanisms.
Therefore, the intuitive conclusion is that reality is more intricate than it might appear at first sight. Furthermore, uncertainties and controversies about the etiological mechanisms driving human neurodegenerative diseases help confound the picture, as do other problems that we will briefly discuss below.
(i) Are inflammasomes functionally interchangeable? Hitherto the interplays that might occur between or among the distinct inflammasomes expressed by each human neural cell type remain mostly undefined. Yet, it is necessary to clarify them to better assess the therapeutic impact of NLRP3 inflammasome inhibitors. Denes et al.’s [336] study results in mice called for caution, as they showed that inflammasomes (e.g., AIM2) can functionally overtake a blocked NLRP3 (Figure 1). A (partial) solution to this problem might entail targeting the ASC protein, which would hinder the activation of all canonical inflammasomes instead of those of NLRP3s only [475]. The inflammasomes’ noncanonical activation problem will persist but might be a minor one.
(ii) The species difference problem. Significant genomic differences apart, not all organs of humans and mammals are morpho-functionally alike. Acceptable similarities exist with liver, kidneys, and lungs. Yet, considering the CNS, while the human cerebral cortex consists mostly of a non-olfactory six-layered neocortex, the widely used rodent models have a less developed, structurally simpler, and mostly olfactory cortex. Moreover, fundamental cytological divergences in size, shape, connections, and functions distinguish the diverse types of neural cells of the human cortex from their rodent counterparts [476]. Human brain’s molecular regulatory mechanisms, e.g., those involved in receptor signal transduction [133] and inflammasome regulation [24,27,477] (see also Box 1 and Box 2), also remarkably diverge from those of rodents. Moreover, human neurodegenerative diseases do not plague rodents in nature. Importantly, in rodent models of human neurodegenerative diseases, the astrocytes undergo an early death—which justifies the often-little attention paid to them—while neurons keep surviving. Conversely, human neurodegenerative diseases kill neurons first, while astrocytes survive and help advance the neuropathologies. Hence, a tight genomic, proteomic, and bio-pathological conformity between animal and human brains is lacking [478,479]. Although brilliant and highly praiseworthy, the manifold animal models of human neurodegenerative diseases in existence cannot surmount such inter-species differences [480]. A quite low animal-to-human translation rate of brain disease-targeting drugs has been persisting for decades, being ascribed to preclinical studies’ faults in “internal consistency” (e.g., design flaws, uncontrolled bias) and/or “external consistency (i.e., animal models pre-testing). As a long trail of clinical trial failures shows, it is difficult to safely predict the effectiveness in humans of drugs pre-tested with favorable results in transgenic animal models [481]. Procedures involving animal models were necessary when nothing or truly little was known about human brain diseases. Now we know much more, albeit not yet enough. Moreover, in recent decades, the legislative/bureaucratic requirements to evaluate novel drugs have become increasingly burdensome to hinder the use of inadequately tested therapeutics. This trend has become stronger after rare events in which properly approved drugs unexpectedly elicited adverse reactions in the patients [482]. Moreover, the repurposing for neurodegenerative diseases of drugs previously evaluated for other ailments in clinical trials is not so easy to do, which precludes the faster testing of potentially useful drugs [483]. Hence, it would be wise to introduce some procedural changes. Animal and/or in silico studies should still help preselect lead drugs. Next, preclinical human untransformed neural cell models in vitro would allow for the assessment of the latter [24,141,212,484] prior to any clinical trial assessment. On rare occasions, animal studies might even be skipped in favor of preclinical human model studies [24,141,212,484]. Human neural cells models will help clarify specific etiopathogenetic mechanisms while supplying safer predicting information about effective drug benefits in clinical settings.
(iii) Symptomatic and/or etiologic therapies? Hitherto, no causal “brain disease modifying” therapies are available for human neurodegenerative diseases. An exception may be the just reported promising effects of Lecanemab, a humanized IgG1 monoclonal antibody binding soluble Aβ protofibrils. After 18 months, Lecanemab reduced brain amyloidosis and slowed cognition decline in early-stage AD patients vs. the placebo-given group. However, Lecanemab also caused collateral brain swelling and/or hemorrhage in some patients, particularly in case of APOE-ε4 homozygotes or anticoagulant therapy [485]. Hence, while Lecanemab’s results confirm that Aβs play a key pathogenetic role in human AD, further studies will prove its etiologic or symptomatic value regarding Aβs/p-Taues’ overproduction and accumulation and inflammasomes’ activity.

6. Conclusions

In recent years, neuroinflammation has been attracting a lot of attention, particularly concerning one of its mediators, i.e., the NLRP3 inflammasome. In the present work, we systematically review the huge and still mounting evidence related to both NLRP3’s involvement in human and animal models of acute and chronic brain diseases, and its many functional activators and inhibitors so far known. Unquestionably, no field expert should disregard the NLRP3 inflammasome, as it is intensely expressed by microglia and circulating monocytes. However, here we wish to stress the indisputable fact that human and animal neural cells of all types, whose morphologies and functions significantly diverge, also express many other inflammasomes and various "anti-inflammasomes"—the latter being tasked with mitigating neuroinflammation. Moreover, the so-called primary drivers of the distinct brain diseases should also be taken into due account because they can simultaneously trigger neurotoxicity and neuroinflammation. Hence, a more comprehensive view of the underlying molecular mechanisms of each brain disease would be beneficial. Importantly, the yet available data on the several inflammasomes’ roles in human brain diseases are limited and controversial. Therefore, this is a field widely open to groundbreaking investigations. We are confident that choosing human untransformed neural cells as models for pathogenetic and pharmacological studies will advance our knowledge about each neuropathology and hasten the achievement of effective etiological therapies.

Author Contributions

Conceptualization, A.C. and U.A.; data curation and investigation, A.C., L.G., C.V., U.A. and I.D.P.; writing—original draft preparation, A.C., L.G., C.V., U.A. and I.D.P.; writing—review and editing, A.C., L.G., C.V., U.A. and I.D.P.; supervision, A.C., L.G., C.V., U.A. and I.D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the FUR 2020 allotment from the Italian MUR (Ministry of University & Research) to A.C. and I.D.P. No funds were provided by private or commercial sources.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors want to thank Saqib Waheed for his expert help with the graphic materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization endorses global action plan on rising incidence of dementia. Nurs. Older People 2017, 29, 7. [CrossRef]
  2. Brett, B.L.; Gardner, R.C.; Godbout, J.; Dams-O′Connor, K.; Keene, C.D. Traumatic Brain Injury and Risk of Neurodegenerative Disorder. Biol. Psychiatry 2022, 91, 498–507. [Google Scholar] [CrossRef] [PubMed]
  3. Walsh, J.G.; Muruve, D.A.; Power, C. Inflammasomes in the CNS. Nat. Rev. Neurosci. 2014, 15, 84–97. [Google Scholar] [CrossRef] [PubMed]
  4. Guo, H.; Callaway, J.B.; Ting, J.P.Y. Inflammasomes: Mechanism of action, role in disease, and therapeutics. Nat. Med. 2015, 21, 677–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Celsius, A.C. De Medicina, Volume 3, Passim; Spencer WG Loeb Classical Library, Translator; Harvard University Press: Cambridge, MA, USA, 1935; ISBN 978-067-499-370-9. [Google Scholar]
  6. Dugger, B.N.; Dickson, D.W. Pathology of Neurodegenerative Diseases. Cold Spring Harb. Perspect. Biol. 2017, 9, a028035. [Google Scholar] [CrossRef] [Green Version]
  7. Wilson, D.M., 3rd; Cookson, M.R.; Van Den Bosch, L.; Zetterberg, H.; Holtzman, D.M.; Dewachter, I. Hallmarks of neurodegenerative diseases. Cell 2023, 186, 693–714. [Google Scholar] [CrossRef]
  8. Tanaka, M.; Toldi, J.; Vécsei, L. Exploring the Etiological Links behind Neurodegenerative Diseases: Inflammatory Cytokines and Bioactive Kynurenines. Int. J. Mol. Sci. 2020, 21, 2431. [Google Scholar] [CrossRef] [Green Version]
  9. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef]
  10. Cao, Z.; Wang, Y.; Long, Z.; He, G. Interaction between autophagy and the NLRP3 inflammasome. Acta Biochim. Biophys. Sin. 2019, 51, 1087–1095. [Google Scholar] [CrossRef]
  11. Bellut, M.; Papp, L.; Bieber, M.; Kraft, P.; Stoll, G.; Schuhmann, M.K. NLPR3 Inflammasome Inhibition Alleviates Hypoxic Endothelial Cell Death in Vitro and Protects Blood–Brain Barrier Integrity in Murine Stroke. Cell Death Dis. 2021, 13, 20. [Google Scholar] [CrossRef]
  12. Kreher, C.; Favret, J.; Maulik, M.; Shin, D. Lysosomal Functions in Glia Associated with Neurodegeneration. Biomolecules 2021, 11, 400. [Google Scholar] [CrossRef]
  13. Chiarini, A.; Dal Pra, I.; Gottardo, R.; Bortolotti, F.; Whitfield, J.F.; Armato, U. BH(4) (tetrahydrobiopterin)-dependent activation, but not the expression, of inducible NOS (nitric oxide synthase)-2 in proinflammatory cytokine-stimulated, cultured normal human astrocytes is mediated by MEK-ERK kinases. J. Cell. Biochem. 2005, 94, 731–743. [Google Scholar] [CrossRef]
  14. Martinon, F.; Burns, K.; Tschopp, J. The inflammasome: A molecular platform triggering activation of inflammatory caspases and processing of proIL-beta. Mol. Cell 2002, 10, 417–426. [Google Scholar] [CrossRef]
  15. de Alba, E. Structure, interactions and self-assembly of ASC-dependent inflammasomes. Arch. Biochem. Biophys. 2019, 670, 15–31. [Google Scholar] [CrossRef]
  16. Stehlik, C.; Lee, S.H.; Dorfleutner, A.; Stassinopoulos, A.; Sagara, J.; Reed, J.C. Apoptosis-associated speck-like protein containing a caspase recruitment domain is a regulator of procaspase-1 activation. J. Immunol. 2003, 171, 6154–6163. [Google Scholar] [CrossRef] [Green Version]
  17. Julien, O.; Wells, J.A. Caspases and their substrates. Cell Death Differ. 2017, 24, 1380–1389. [Google Scholar] [CrossRef]
  18. Ding, J.; Wang, K.; Liu, W.; She, Y.; Sun, Q.; Shi, J.; Sun, H.; Wang, D.C.; Shao, F. Pore-forming activity and structural autoinhibition of the gasdermin family. Nature 2016, 535, 111–116. [Google Scholar] [CrossRef]
  19. Gambin, Y.; Giles, N.; O′Carroll, A.; Polinkovsky, M.; Hunter, D.; Sierecki, E. Single-molecule fluorescence reveals the oligomerization and folding steps driving the prion-like behavior of ASC. J. Mol. Biol. 2018, 430, 491–508. [Google Scholar] [CrossRef] [Green Version]
  20. Kesavardhana, S.; Kanneganti, T.D. Mechanisms governing inflammasome activation, assembly and pyroptosis induction. Int. Immunol. 2017, 29, 201–210. [Google Scholar] [CrossRef] [Green Version]
  21. Lupfer, C.; Kanneganti, T.D. Unsolved mysteries in NLR biology. Front. Immunol. 2013, 4, 285. [Google Scholar] [CrossRef] [Green Version]
  22. Devi, S.; Stehlik, C.; Dorfleutner, A. An update on CARD only proteins (COPs) and PYD only proteins (POPs) as inflammasome regulators. Int. J. Mol. Sci. 2020, 21, 6901. [Google Scholar] [CrossRef] [PubMed]
  23. Poli, G.; Fabi, C.; Bellet, M.M.; Costantini, C.; Nunziangeli, L.; Romani, L.; Brancorsini, S. Epigenetic mechanisms of inflammasome regulation. Int. J. Mol. Sci. 2020, 21, 5758. [Google Scholar] [CrossRef] [PubMed]
  24. Chiarini, A.; Armato, U.; Gui, L.; Dal Prà, I. “Other Than NLRP3” Inflammasomes: Multiple Roles in Brain Disease. Neuroscientist 2022, 11, 10738584221106114. [Google Scholar] [CrossRef] [PubMed]
  25. Mangan, M.S.J.; Olhava, E.J.; Roush, W.R.; Seidel, H.M.; Glick, G.D.; Latz, E. Targeting the NLRP3 inflammasome in inflammatory diseases. Nat. Rev. Drug Discov. 2018, 17, 588–606. [Google Scholar] [CrossRef]
  26. Chen, J.; Chen, Z.J. PtdIns4P on dispersed trans-Golgi network mediates NLRP3 inflammasome activation. Nature 2018, 564, 71–76. [Google Scholar] [CrossRef]
  27. Chiarini, A.; Armato, U.; Hu, P.; Dal Prà, I. Danger-sensing/pattern recognition receptors and neuroinflammation in Alzheimer′s disease. Int. J. Mol. Sci. 2020, 21, 9036. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Zhao, Y.; Zhang, J.; Yang, G. Mechanisms of NLRP3 Inflammasome Activation: Its Role in the Treatment of Alzheimer′s Disease. Neurochem. Res. 2020, 45, 2560–2572. [Google Scholar] [CrossRef]
  29. Holbrook, J.A.; Jarosz-Griffiths, H.H.; Caseley, E.; Lara-Reyna, S.; Poulter, J.A.; Williams-Gray, C.H.; Peckham, D.; McDermott, M.F. Neurodegenerative Disease and the NLRP3 Inflammasome. Front. Pharmacol. 2021, 12, 643254. [Google Scholar] [CrossRef]
  30. Mészáros, Á.; Molnár, K.; Nógrádi, B.; Hernádi, Z.; Nyúl-Tóth, Á.; Wilhelm, I.; Krizbai, I.A. Neurovascular Inflammaging in Health and Disease. Cells 2020, 9, 1614. [Google Scholar] [CrossRef]
  31. Lee, G.S.; Subramanian, N.; Kim, A.I.; Aksentijevich, I.; Goldbach-Mansky, R.; Sacks, D.B.; Germain, R.N.; Kastner, D.L.; Chae, J.J. The calcium-sensing receptor regulates the NLRP3 inflammasome through Ca2+ and cAMP. Nature 2012, 492, 123–127. [Google Scholar] [CrossRef] [Green Version]
  32. Gong, Z.; Pan, J.; Shen, Q.; Li, M.; Peng, Y. Mitochondrial dysfunction induces NLRP3 inflammasome activation during cerebral ischemia/reperfusion injury. J. Neuroinflamm. 2018, 15, 242. [Google Scholar] [CrossRef] [Green Version]
  33. Su, S.H.; Wu, Y.F.; Wang, D.P.; Hai, J. Inhibition of excessive autophagy and mitophagy mediates neuroprotective effects of URB597 against chronic cerebral hypoperfusion. Cell Death Dis. 2018, 9, 733. [Google Scholar] [CrossRef] [Green Version]
  34. Su, S.H.; Wu, Y.F.; Lin, Q.; Wang, D.P.; Hai, J. URB597 protects against NLRP3 inflammasome activation by inhibiting autophagy dysfunction in a rat model of chronic cerebral hypoperfusion. J. Neuroinflamm. 2019, 16, 260. [Google Scholar] [CrossRef] [Green Version]
  35. Zhu, J.J.; Yu, B.Y.; Huang, X.K.; He, M.Z.; Chen, B.W.; Chen, T.T.; Fang, H.Y.; Chen, S.Q.; Fu, X.Q.; Li, P.J.; et al. Neferine Protects against Hypoxic-Ischemic Brain Damage in Neonatal Rats by Suppressing NLRP3-Mediated Inflammasome Activation. Oxid. Med. Cell. Longev. 2021, 2021, 6654954. [Google Scholar] [CrossRef]
  36. Franke, M.; Bieber, M.; Kraft, P.; Weber, A.N.R.; Stoll, G.; Schuhmann, M.K. The NLRP3 inflammasome drives inflammation in ischemia/reperfusion injury after transient middle cerebral artery occlusion in mice. Brain Behav. Immun. 2021, 92, 223–233. [Google Scholar] [CrossRef]
  37. Xu, Q.; Zhao, B.; Ye, Y.; Li, Y.; Zhang, Y.; Xiong, X.; Gu, L. Relevant mediators involved in and therapies targeting the inflammatory response induced by activation of the NLRP3 inflammasome in ischemic stroke. J. Neuroinflamm. 2021, 18, 123. [Google Scholar] [CrossRef]
  38. Chen, S.H.; Scott, X.O.; Ferrer Marcelo, Y.; Almeida, V.W.; Blackwelder, P.L.; Yavagal, D.R.; Peterson, E.C.; Starke, R.M.; Dietrich, W.D.; Keane, R.W.; et al. Netosis and Inflammasomes in Large Vessel Occlusion Thrombi. Front. Pharmacol. 2021, 11, 607287. [Google Scholar] [CrossRef]
  39. Xiao, L.; Zheng, H.; Li, J.; Wang, Q.; Sun, H. Neuroinflammation Mediated by NLRP3 Inflammasome after Intracerebral Hemorrhage and Potential Therapeutic Targets. Mol. Neurobiol. 2020, 57, 5130–5149. [Google Scholar] [CrossRef]
  40. Yang, S.J.; Shao, G.F.; Chen, J.L.; Gong, J. The NLRP3 Inflammasome: An Important Driver of Neuroinflammation in Hemorrhagic Stroke. Cell. Mol. Neurobiol. 2018, 38, 595–603. [Google Scholar] [CrossRef]
  41. Cristina de Brito Toscano, E.; Leandro Marciano Vieira, É.; Boni Rocha Dias, B.; Vidigal Caliari, M.; Paula Gonçalves, A.; Varela Giannetti, A.; Maurício Siqueira, J.; Kimie Suemoto, C.; Elaine Paraizo Leite, R.; Nitrini, R.; et al. NLRP3 and NLRP1 inflammasomes are up-regulated in patients with mesial temporal lobe epilepsy and may contribute to overexpression of caspase-1 and IL-β in sclerotic hippocampi. Brain Res. 2021, 1752, 147230. [Google Scholar] [CrossRef]
  42. Wang, S.; He, H.; Long, J.; Sui, X.; Yang, J.; Lin, G.; Wang, Q.; Wang, Y.; Luo, Y. TRPV4 Regulates Soman-Induced Status Epilepticus and Secondary Brain Injury via NMDA Receptor and NLRP3 Inflammasome. Neurosci. Bull. 2021, 37, 905–920. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, D.; Zhang, J.; Jiang, W.; Cao, Z.; Zhao, F.; Cai, T.; Aschner, M.; Luo, W. The role of NLRP3-CASP1 in inflammasome-mediated neuroinflammation and autophagy dysfunction in manganese-induced, hippocampal-dependent impairment of learning and memory ability. Autophagy 2017, 13, 914–927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Sarkar, S.; Rokad, D.; Malovic, E.; Luo, J.; Harischandra, D.S.; Jin, H.; Anantharam, V.; Huang, X.; Lewis, M.; Kanthasamy, A.; et al. Manganese activates NLRP3 inflammasome signaling and propagates exosomal release of ASC in microglial cells. Sci. Signal. 2019, 12, eaat9900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Su, P.; Wang, D.; Cao, Z.; Chen, J.; Zhang, J. The role of NLRP3 in lead-induced neuroinflammation and possible underlying mechanism. Environ. Pollut. 2021, 287, 117520. [Google Scholar] [CrossRef] [PubMed]
  46. Dong, J.; Wang, X.; Xu, C.; Gao, M.; Wang, S.; Zhang, J.; Tong, H.; Wang, L.; Han, Y.; Cheng, N.; et al. Inhibiting NLRP3 inflammasome activation prevents copper-induced neuropathology in a murine model of Wilson′s disease. Cell Death Dis. 2021, 12, 87. [Google Scholar] [CrossRef]
  47. Cai, J.; Guan, H.; Jiao, X.; Yang, J.; Chen, X.; Zhang, H.; Zheng, Y.; Zhu, Y.; Liu, Q.; Zhang, Z. NLRP3 inflammasome mediated pyroptosis is involved in cadmium exposure-induced neuroinflammation through the IL-1β/IkB-α-NF-κB-NLRP3 feedback loop in swine. Toxicology 2021, 453, 152720. [Google Scholar] [CrossRef]
  48. Brewer, G.J. Divalent Copper as a Major Triggering Agent in Alzheimer′s Disease. J. Alzheimer’s Dis. 2015, 46, 593–604. [Google Scholar] [CrossRef]
  49. Zhou, Q.; Zhang, Y.; Lu, L.; Zhang, H.; Zhao, C.; Pu, Y.; Yin, L. Copper induces microglia-mediated neuroinflammation through ROS/NF-κB pathway and mitophagy disorder. Food Chem. Toxicol. 2022, 16, 113369. [Google Scholar] [CrossRef]
  50. Quandt, D.; Rothe, K.; Baerwald, C.; Rossol, M. GPRC6A mediates Alum-induced Nlrp3 inflammasome activation but limits Th2 type antibody responses. Sci. Rep. 2015, 5, 16719. [Google Scholar] [CrossRef] [Green Version]
  51. Ye, R.; Pi, M.; Nooh, M.M.; Bahout, S.W.; Quarles, L.D. Human GPRC6A Mediates Testosterone-Induced Mitogen-Activated Protein Kinases and mTORC1 Signaling in Prostate Cancer Cells. Mol. Pharmacol. 2019, 95, 563–572. [Google Scholar] [CrossRef] [Green Version]
  52. Zhang, X.; Shu, Q.; Liu, Z.; Gao, C.; Wang, Z.; Xing, Z.; Song, J. Recombinant osteopontin provides protection for cerebral infarction by inhibiting the NLRP3 inflammasome in microglia. Brain Res. 2021, 1751, 147170. [Google Scholar] [CrossRef]
  53. Chen, Y.; Meng, J.; Bi, F.; Li, H.; Chang, C.; Ji, C.; Liu, W. NEK7 Regulates NLRP3 Inflammasome Activation and Neuroinflammation Post-traumatic Brain Injury. Front. Mol. Neurosci. 2019, 12, 202, Erratum in Front. Mol. Neurosci. 2019, 12, 247. [Google Scholar] [CrossRef] [Green Version]
  54. Ji, X.; Song, Z.; He, J.; Guo, S.; Chen, Y.; Wang, H.; Zhang, J.; Xu, X.; Liu, J. NIMA-related kinase 7 amplifies NLRP3 inflammasome pro-inflammatory signaling in microglia/macrophages and mice models of spinal cord injury. Exp. Cell Res. 2021, 398, 112418. [Google Scholar] [CrossRef]
  55. O′Brien, W.T.; Pham, L.; Symons, G.F.; Monif, M.; Shultz, S.R.; McDonald, S.J. The NLRP3 inflammasome in traumatic brain injury: Potential as a biomarker and therapeutic target. J. Neuroinflamm. 2020, 17, 104. [Google Scholar] [CrossRef]
  56. Irrera, N.; Russo, M.; Pallio, G.; Bitto, A.; Mannino, F.; Minutoli, L.; Altavilla, D.; Squadrito, F. The Role of NLRP3 Inflammasome in the Pathogenesis of Traumatic Brain Injury. Int. J. Mol. Sci. 2020, 21, 6204. [Google Scholar] [CrossRef]
  57. Albalawi, F.; Lu, W.; Beckel, J.M.; Lim, J.C.; McCaughey, S.A.; Mitchell, C.H. The P2X7 Receptor Primes IL-1β and the NLRP3 Inflammasome in Astrocytes Exposed to Mechanical Strain. Front. Cell. Neurosci. 2017, 11, 227. [Google Scholar] [CrossRef]
  58. Ding, H.; Li, Y.; Wen, M.; Liu, X.; Han, Y.; Zeng, H. Elevated intracranial pressure induces IL1β and IL18 overproduction via activation of the NLRP3 inflammasome in microglia of ischemic adult rats. Int. J. Mol. Med. 2021, 47, 183–194. [Google Scholar] [CrossRef]
  59. Chi, W.; Chen, H.; Li, F.; Zhu, Y.; Yin, W.; Zhuo, Y. HMGB1 promotes the activation of NLRP3 and caspase-8 inflammasomes via NF-κB pathway in acute glaucoma. J. Neuroinflamm. 2015, 12, 137. [Google Scholar] [CrossRef] [Green Version]
  60. Jiang, S.; Maphis, N.M.; Binder, J.; Chisholm, D.; Weston, L.; Duran, W.; Peterson, C.; Zimmerman, A.; Mandell, M.A.; Jett, S.D.; et al. Proteopathic tau primes and activates interleukin-1β via myeloid-cell-specific MyD88- and NLRP3-ASC-inflammasome pathway. Cell Rep. 2021, 36, 109720. [Google Scholar] [CrossRef]
  61. Shi, F.; Yang, L.; Kouadir, M.; Yang, Y.; Wang, J.; Zhou, X.; Yin, X.; Zhao, D. The NALP3 inflammasome engages in neurotoxic prion peptide-induced microglial activation. J. Neuroinflamm. 2012, 9, 73. [Google Scholar] [CrossRef] [Green Version]
  62. Lai, M.; Yao, H.; Shah, S.Z.A.; Wu, W.; Wang, D.; Zhao, Y.; Wang, L.; Zhou, X.; Zhao, D.; Yang, L. The NLRP3-Caspase 1 Inflammasome Negatively Regulates Autophagy via TLR4-TRIF in Prion Peptide-Infected Microglia. Front. Aging Neurosci. 2018, 10, 116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Milner, M.T.; Maddugoda, M.; Götz, J.; Burgener, S.S.; Schroder, K. The NLRP3 inflammasome triggers sterile neuroinflammation and Alzheimer′s disease. Curr. Opin. Immunol. 2021, 68, 116–124. [Google Scholar] [CrossRef] [PubMed]
  64. Pike, A.F.; Varanita, T.; Herrebout, M.A.C.; Plug, B.C.; Kole, J.; Musters, R.J.P.; Teunissen, C.E.; Hoozemans, J.J.M.; Bubacco, L.; Veerhuis, R. α-Synuclein evokes NLRP3 inflammasome-mediated IL-1β secretion from primary human microglia. Glia 2021, 69, 1413–1428. [Google Scholar] [CrossRef] [PubMed]
  65. Deora, V.; Lee, J.D.; Albornoz, E.A.; McAlary, L.; Jagaraj, C.J.; Robertson, A.A.B.; Atkin, J.D.; Cooper, M.A.; Schroder, K.; Yerbury, J.J.; et al. The microglial NLRP3 inflammasome is activated by amyotrophic lateral sclerosis proteins. Glia 2020, 68, 407–421, Erratum in Glia 2020, 68, 2167–2168. [Google Scholar] [CrossRef] [PubMed]
  66. Ismael, S.; Nasoohi, S.; Li, L.; Aslam, K.S.; Khan, M.M.; El-Remessy, A.B.; McDonald, M.P.; Liao, F.F.; Ishrat, T. Thioredoxin interacting protein regulates age-associated neuroinflammation. Neurobiol. Dis. 2021, 156, 105399. [Google Scholar] [CrossRef]
  67. Ismael, S.; Wajidunnisa; Sakata, K.; McDonald, M.P.; Liao, F.F.; Ishrat, T. ER stress associated TXNIP-NLRP3 inflammasome activation in hippocampus of human Alzheimer′s disease. Neurochem. Int. 2021, 148, 105104. [Google Scholar] [CrossRef]
  68. Shen, H.; Guan, Q.; Zhang, X.; Yuan, C.; Tan, Z.; Zhai, L.; Hao, Y.; Gu, Y.; Han, C. New mechanism of neuroinflammation in Alzheimer′s disease: The activation of NLRP3 inflammasome mediated by gut microbiota. Prog. Neuropsychopharmacol. Biol. Psychiatry 2020, 100, 109884. [Google Scholar] [CrossRef]
  69. Shukla, P.K.; Delotterie, D.F.; Xiao, J.; Pierre, J.F.; Rao, R.; McDonald, M.P.; Khan, M.M. Alterations in the Gut-Microbial-Inflammasome-Brain Axis in a Mouse Model of Alzheimer′s Disease. Cells 2021, 10, 779. [Google Scholar] [CrossRef]
  70. Yi, W.; Cheng, J.; Wei, Q.; Pan, R.; Song, S.; He, Y.; Tang, C.; Liu, X.; Zhou, Y.; Su, H. Effect of temperature stress on gut-brain axis in mice: Regulation of intestinal microbiome and central NLRP3 inflammasomes. Sci. Total Environ. 2021, 772, 144568. [Google Scholar] [CrossRef]
  71. Ising, C.; Venegas, C.; Zhang, S.; Scheiblich, H.; Schmidt, S.V.; Vieira-Saecker, A.; Schwartz, S.; Albasset, S.; McManus, R.M.; Tejera, D.; et al. NLRP3 inflammasome activation drives tau pathology. Nature 2019, 575, 669–673. [Google Scholar] [CrossRef]
  72. Wang, B.R.; Shi, J.Q.; Ge, N.N.; Ou, Z.; Tian, Y.Y.; Jiang, T.; Zhou, J.S.; Xu, J.; Zhang, Y.D. PM2.5 exposure aggravates oligomeric amyloid beta-induced neuronal injury and promotes NLRP3 inflammasome activation in an in vitro model of Alzheimer′s disease. J. Neuroinflamm. 2018, 15, 132. [Google Scholar] [CrossRef] [Green Version]
  73. Shi, J.Q.; Wang, B.R.; Jiang, T.; Gao, L.; Zhang, Y.D.; Xu, J. NLRP3 Inflammasome: A Potential Therapeutic Target in Fine Particulate Matter-Induced Neuroinflammation in Alzheimer′s Disease. J. Alzheimers Dis. 2020, 77, 923–934. [Google Scholar] [CrossRef]
  74. Yuan, L.; Zhu, Y.; Huang, S.; Lin, L.; Jiang, X.; Chen, S. NF-κB/ROS and ERK pathways regulate NLRP3 inflammasome activation in Listeria monocytogenes infected BV2 microglia cells. J. Microbiol. 2021, 59, 771–781. [Google Scholar] [CrossRef]
  75. Zhao, Z.; Wang, Y.; Zhou, R.; Li, Y.; Gao, Y.; Tu, D.; Wilson, B.; Song, S.; Feng, J.; Hong, J.S.; et al. A novel role of NLRP3-generated IL-1β in the acute-chronic transition of peripheral lipopolysaccharide-elicited neuroinflammation: Implications for sepsis-associated neurodegeneration. J. Neuroinflamm. 2020, 17, 64. [Google Scholar] [CrossRef] [Green Version]
  76. Danielski, L.G.; Giustina, A.D.; Bonfante, S.; de Souza Goldim, M.P.; Joaquim, L.; Metzker, K.L.; Biehl, E.B.; Vieira, T.; de Medeiros, F.D.; da Rosa, N.; et al. NLRP3 Activation Contributes to Acute Brain Damage Leading to Memory Impairment in Sepsis-Surviving Rats. Mol. Neurobiol. 2020, 57, 5247–5262. [Google Scholar] [CrossRef]
  77. Chivero, E.T.; Guo, M.L.; Periyasamy, P.; Liao, K.; Callen, S.E.; Buch, S. HIV-1 Tat Primes and Activates Microglial NLRP3 Inflammasome-Mediated Neuroinflammation. J. Neurosci. 2017, 37, 3599–3609. [Google Scholar] [CrossRef] [Green Version]
  78. Katuri, A.; Bryant, J.; Heredia, A.; Makar, T.K. Role of the inflammasomes in HIV-associated neuroinflammation and neurocognitive disorders. Exp. Mol. Pathol. 2019, 108, 64–72. [Google Scholar] [CrossRef]
  79. He, X.; Yang, W.; Zeng, Z.; Wei, Y.; Gao, J.; Zhang, B.; Li, L.; Liu, L.; Wan, Y.; Zeng, Q.; et al. NLRP3-dependent pyroptosis is required for HIV-1 gp120-induced neuropathology. Cell. Mol. Immunol. 2020, 17, 283–299. [Google Scholar] [CrossRef]
  80. Hu, X.; Zeng, Q.; Xiao, J.; Qin, S.; Wang, Y.; Shan, T.; Hu, D.; Zhu, Y.; Liu, K.; Zheng, K.; et al. Herpes Simplex Virus 1 Induces Microglia Gasdermin D-Dependent Pyroptosis through Activating the NLR Family Pyrin Domain Containing 3 Inflammasome. Front. Microbiol. 2022, 13, 838808. [Google Scholar] [CrossRef]
  81. Chen, C.J.; Ou, Y.C.; Chang, C.Y.; Pan, H.C.; Lin, S.Y.; Liao, S.L.; Raung, S.L.; Chen, S.Y.; Chang, C.J. Src signaling involvement in Japanese encephalitis virus-induced cytokine production in microglia. Neurochem. Int. 2011, 58, 924–933. [Google Scholar] [CrossRef]
  82. He, Z.; Chen, J.; Zhu, X.; An, S.; Dong, X.; Yu, J.; Zhang, S.; Wu, Y.; Li, G.; Zhang, Y.; et al. NLRP3 Inflammasome Activation Mediates Zika Virus-Associated Inflammation. J. Infect. Dis. 2018, 217, 1942–1951. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, I.Y.; Moriyama, M.; Chang, M.F.; Ichinohe, T. Severe Acute Respiratory Syndrome Coronavirus Viroporin 3a Activates the NLRP3 Inflammasome. Front. Microbiol. 2019, 10, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Siu, K.L.; Yuen, K.S.; Castano-Rodriguez, C.; Ye, Z.W.; Yeung, M.L.; Fung, S.Y.; Yuan, S.; Chan, C.P.; Yuen, K.Y.; Enjuanes, L.; et al. Severe acute respiratory syndrome coronavirus ORF3a protein activates the NLRP3 inflammasome by promoting TRAF3-dependent ubiquitination of ASC. FASEB J. 2019, 33, 8865–8877. [Google Scholar] [CrossRef] [PubMed]
  85. de Rivero Vaccari, J.C.; Dietrich, W.D.; Keane, R.W.; de Rivero Vaccari, J.P. The inflammasome in times of COVID-19. Front. Immunol. 2020, 11, 583373. [Google Scholar] [CrossRef]
  86. Pan, P.; Shen, M.; Yu, Z.; Ge, W.; Chen, K.; Tian, M.; Xiao, F.; Wang, Z.; Wang, J.; Jia, Y.; et al. SARS-CoV-2 N protein promotes NLRP3 inflammasome activation to induce hyperinflammation. Nat. Commun. 2021, 12, 4664, Erratum in Nat. Commun. 2021, 12, 5306. [Google Scholar] [CrossRef]
  87. Yalcinkaya, M.; Liu, W.; Islam, M.N.; Kotini, A.G.; Gusarova, G.A.; Fidler, T.P.; Papapetrou, E.P.; Bhattacharya, J.; Wang, N.; Tall, A.R. Modulation of the NLRP3 inflammasome by SARS-CoV-2 Envelope protein. Sci. Rep. 2021, 11, 24432. [Google Scholar] [CrossRef]
  88. Olajide, O.A.; Iwuanyanwu, V.U.; Adegbola, O.D.; Al-Hindawi, A.A. SARS-CoV-2 Spike Glycoprotein S1 Induces Neuroinflammation in BV-2 Microglia. Mol. Neurobiol. 2022, 59, 445–458. [Google Scholar] [CrossRef]
  89. Ito, M.; Yanagi, Y.; Ichinohe, T. Encephalomyocarditis virus viroporin 2B activates NLRP3 inflammasome. PLoS Pathog. 2012, 8, e1002857. [Google Scholar] [CrossRef] [Green Version]
  90. Moreira, J.D.; Iakhiaev, A.; Vankayalapati, R.; Jung, B.G.; Samten, B. Histone deacetylase-2 controls IL-1β production through the regulation of NLRP3 expression and activation in tuberculosis infection. iScience 2022, 25, 104799. [Google Scholar] [CrossRef]
  91. Butterfield, D.A.; Halliwell, B. Oxidative stress, dysfunctional glucose metabolism and Alzheimer disease. Nat. Rev. Neurosci. 2019, 20, 148–160. [Google Scholar] [CrossRef]
  92. Litwiniuk, A.; Bik, W.; Kalisz, M.; Baranowska-Bik, A. Inflammasome NLRP3 Potentially Links Obesity-Associated Low-Grade Systemic Inflammation and Insulin Resistance with Alzheimer′s Disease. Int. J. Mol. Sci. 2021, 22, 5603. [Google Scholar] [CrossRef]
  93. Sobesky, J.L.; D′Angelo, H.M.; Weber, M.D.; Anderson, N.D.; Frank, M.G.; Watkins, L.R.; Maier, S.F.; Barrientos, R.M. Glucocorticoids Mediate Short-Term High-Fat Diet Induction of Neuroinflammatory Priming, the NLRP3 Inflammasome, and the Danger Signal HMGB1. eNeuro 2016, 3, ENEURO.0113-16.2016. [Google Scholar] [CrossRef] [Green Version]
  94. Keshk, W.A.; Ibrahim, M.A.; Shalaby, S.M.; Zalat, Z.A.; Elseady, W.S. Redox status, inflammation, necroptosis and inflammasome as indispensable contributors to high fat diet (HFD)-induced neurodegeneration; Effect of N-acetylcysteine (NAC). Arch. Biochem. Biophys. 2020, 680, 108227. [Google Scholar] [CrossRef]
  95. Wei, P.; Yang, F.; Zheng, Q.; Tang, W.; Li, J. The Potential Role of the NLRP3 Inflammasome Activation as a Link between Mitochondria ROS Generation and Neuroinflammation in Postoperative Cognitive Dysfunction. Front. Cell. Neurosci. 2019, 13, 73. [Google Scholar] [CrossRef] [Green Version]
  96. Zhang, L.; Xiao, F.; Zhang, J.; Wang, X.; Ying, J.; Wei, G.; Chen, S.; Huang, X.; Yu, W.; Liu, X.; et al. Dexmedetomidine Mitigated NLRP3-Mediated Neuroinflammation via the Ubiquitin-Autophagy Pathway to Improve Perioperative Neurocognitive Disorder in Mice. Front. Pharmacol. 2021, 12, 646265. [Google Scholar] [CrossRef]
  97. Hirshman, N.A.; Hughes, F.M., Jr.; Jin, H.; Harrison, W.T.; White, S.W.; Doan, I.; Harper, S.N.; Leidig, P.D.; Purves, J.T. Cyclophosphamide-induced cystitis results in NLRP3-mediated inflammation in the hippocampus and symptoms of depression in rats. Am. J. Physiol. Renal Physiol. 2020, 318, F354–F362. [Google Scholar] [CrossRef]
  98. D′Espessailles, A.; Mora, Y.A.; Fuentes, C.; Cifuentes, M. Calcium-sensing receptor activates the NLRP3 inflammasome in LS14 preadipocytes mediated by ERK1/2 signaling. J. Cell. Physiol. 2018, 233, 6232–6240. [Google Scholar] [CrossRef]
  99. Wang, C.; Jia, Q.; Sun, C.; Jing, C. Calcium sensing receptor contribute to early brain injury through the CaMKII/NLRP3 pathway after subarachnoid hemorrhage in mice. Biochem. Biophys. Res. Commun. 2020, 530, 651–657. [Google Scholar] [CrossRef]
  100. Hu, W.; Zhang, Y.; Wu, W.; Yin, Y.; Huang, D.; Wang, Y.; Li, W.; Li, W. Chronic glucocorticoids exposure enhances neurodegeneration in the frontal cortex and hippocampus via NLRP-1 inflammasome activation in male mice. Brain Behav. Immun. 2016, 52, 58–70. [Google Scholar] [CrossRef]
  101. Maturana, C.J.; Aguirre, A.; Sáez, J.C. High glucocorticoid levels during gestation activate the inflammasome in hippocampal oligodendrocytes of the offspring. Dev. Neurobiol. 2017, 77, 625–642. [Google Scholar] [CrossRef]
  102. Chivero, E.T.; Thangaraj, A.; Tripathi, A.; Periyasamy, P.; Guo, M.L.; Buch, S. NLRP3 Inflammasome Blockade Reduces Cocaine-Induced Microglial Activation and Neuroinflammation. Mol. Neurobiol. 2021, 58, 2215–2230. [Google Scholar] [CrossRef] [PubMed]
  103. Du, S.H.; Qiao, D.F.; Chen, C.X.; Chen, S.; Liu, C.; Lin, Z.; Wang, H.; Xie, W.B. Toll-Like Receptor 4 Mediates Methamphetamine-Induced Neuroinflammation through Caspase-11 Signaling Pathway in Astrocytes. Front. Mol. Neurosci. 2017, 10, 409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Xu, E.; Liu, J.; Liu, H.; Wang, X.; Xiong, H. Inflammasome Activation by Methamphetamine Potentiates Lipopolysaccharide Stimulation of IL-1β Production in Microglia. J. Neuroimmune Pharmacol. 2018, 13, 237–253. [Google Scholar] [CrossRef] [PubMed]
  105. Cheon, S.Y.; Koo, B.N.; Kim, S.Y.; Kam, E.H.; Nam, J.; Kim, E.J. Scopolamine promotes neuroinflammation and delirium-like neuropsychiatric disorder in mice. Sci. Rep. 2021, 11, 8376. [Google Scholar] [CrossRef]
  106. Lippai, D.; Bala, S.; Petrasek, J.; Csak, T.; Levin, I.; Kurt-Jones, E.A.; Szabo, G. Alcohol-induced IL-1β in the brain is mediated by NLRP3/ASC inflammasome activation that amplifies neuroinflammation. J. Leukoc. Biol. 2013, 94, 171–182. [Google Scholar] [CrossRef] [Green Version]
  107. Alfonso-Loeches, S.; Ureña-Peralta, J.; Morillo-Bargues, M.J.; Gómez-Pinedo, U.; Guerri, C. Ethanol-Induced TLR4/NLRP3 Neuroinflammatory Response in Microglial Cells Promotes Leukocyte Infiltration Across the BBB. Neurochem. Res. 2016, 41, 193–209. [Google Scholar] [CrossRef]
  108. Carranza-Aguilar, C.J.; Hernández-Mendoza, A.; Mejias-Aponte, C.; Rice, K.C.; Morales, M.; González-Espinosa, C.; Cruz, S.L. Morphine and Fentanyl Repeated Administration Induces Different Levels of NLRP3-Dependent Pyroptosis in the Dorsal Raphe Nucleus of Male Rats via Cell-Specific Activation of TLR4 and Opioid Receptors. Cell. Mol. Neurobiol. 2020, 42, 677–694. [Google Scholar] [CrossRef]
  109. Samir, P.; Kesavardhana, S.; Patmore, D.M.; Gingras, S.; Malireddi, R.K.S.; Karki, R.; Guy, C.S.; Briard, B.; Place, D.E.; Bhattacharya, A.; et al. DDX3X acts as a live-or-die checkpoint in stressed cells by regulating NLRP3 inflammasome. Nature 2019, 573, 590–594. [Google Scholar] [CrossRef]
  110. Swaroop, S.; Mahadevan, A.; Shankar, S.K.; Adlakha, Y.K.; Basu, A. HSP60 critically regulates endogenous IL-1β production in activated microglia by stimulating NLRP3 inflammasome pathway. J. Neuroinflamm. 2018, 15, 177, Erratum in J. Neuroinflamm. 2018, 15, 317. [Google Scholar] [CrossRef] [Green Version]
  111. Kim, M.J.; Yoon, J.H.; Ryu, J.H. Mitophagy: A balance regulator of NLRP3 inflammasome activation. BMB Rep. 2016, 49, 529–535. [Google Scholar] [CrossRef] [Green Version]
  112. Mishra, S.R.; Mahapatra, K.K.; Behera, B.P.; Patra, S.; Bhol, C.S.; Panigrahi, D.P.; Praharaj, P.P.; Singh, A.; Patil, S.; Dhiman, R.; et al. Mitochondrial dysfunction as a driver of NLRP3 inflammasome activation and its modulation through mitophagy for potential therapeutics. Int. J. Biochem. Cell Biol. 2021, 136, 106013. [Google Scholar] [CrossRef]
  113. Leemans, J.C.; Cassel, S.L.; Sutterwala, F.S. Sensing damage by the NLRP3 inflammasome. Immunol. Rev. 2011, 243, 152–162. [Google Scholar] [CrossRef] [Green Version]
  114. Iyer, S.S.; He, Q.; Janczy, J.R.; Elliott, E.I.; Zhong, Z.; Olivier, A.K.; Sadler, J.J.; Knepper-Adrian, V.; Han, R.; Qiao, L.; et al. Mitochondrial cardiolipin is required for Nlrp3 inflammasome activation. Immunity 2013, 39, 311–323. [Google Scholar] [CrossRef] [Green Version]
  115. Han, S.; He, Z.; Jacob, C.; Hu, X.; Liang, X.; Xiao, W.; Wan, L.; Xiao, P.; D′Ascenzo, N.; Ni, J.; et al. Effect of Increased IL-1β on Expression of HK in Alzheimer′s Disease. Int. J. Mol. Sci. 2021, 22, 1306. [Google Scholar] [CrossRef]
  116. Rivers-Auty, J.; Tapia, V.S.; White, C.S.; Daniels, M.J.D.; Drinkall, S.; Kennedy, P.T.; Spence, H.G.; Yu, S.; Green, J.P.; Hoyle, C.; et al. Zinc Status Alters Alzheimer′s Disease Progression through NLRP3-Dependent Inflammation. J. Neurosci. 2021, 41, 3025–3038. [Google Scholar] [CrossRef]
  117. Xu, Z.; Chen, Z.M.; Wu, X.; Zhang, L.; Cao, Y.; Zhou, P. Distinct Molecular Mechanisms Underlying Potassium Efflux for NLRP3 Inflammasome Activation. Front. Immunol. 2020, 11, 609441. [Google Scholar] [CrossRef]
  118. Zhong, Z.; Liang, S.; Sanchez-Lopez, E.; He, F.; Shalapour, S.; Lin, X.J.; Wong, J.; Ding, S.; Seki, E.; Schnabl, B.; et al. New mitochondrial DNA synthesis enables NLRP3 inflammasome activation. Nature 2018, 560, 198–203. [Google Scholar] [CrossRef]
  119. Zhou, X.G.; Qiu, W.Q.; Yu, L.; Pan, R.; Teng, J.F.; Sang, Z.P.; Law, B.Y.; Zhao, Y.; Zhang, L.; Yan, L.; et al. Targeting microglial autophagic degradation of the NLRP3 inflammasome for identification of thonningianin A in Alzheimer′s disease. Inflamm. Regen. 2022, 42, 25. [Google Scholar] [CrossRef]
  120. Zhao, T.; Gao, J.; Van, J.; To, E.; Wang, A.; Cao, S.; Cui, J.Z.; Guo, J.P.; Lee, M.; McGeer, P.L.; et al. Age-related increases in amyloid beta and membrane attack complex: Evidence of inflammasome activation in the rodent eye. J. Neuroinflamm. 2015, 12, 121. [Google Scholar] [CrossRef] [Green Version]
  121. Reddy, P.H.; Oliver, D.M. Amyloid Beta and Phosphorylated Tau-Induced Defective Autophagy and Mitophagy in Alzheimer′s Disease. Cells 2019, 8, 488. [Google Scholar] [CrossRef] [Green Version]
  122. Eshraghi, M.; Adlimoghaddam, A.; Mahmoodzadeh, A.; Sharifzad, F.; Yasavoli-Sharahi, H.; Lorzadeh, S.; Albensi, B.C.; Ghavami, S. Alzheimer′s Disease Pathogenesis: Role of Autophagy and Mitophagy Focusing in Microglia. Int. J. Mol. Sci. 2021, 22, 3330. [Google Scholar] [CrossRef] [PubMed]
  123. Lech, M.; Avila-Ferrufino, A.; Skuginna, V.; Susanti, H.E.; Anders, H.J. Quantitative expression of RIG-like helicase, NOD-like receptor and inflammasome-related mRNAs in humans and mice. Int. Immunol. 2010, 22, 717–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Minkiewicz, J.; de Rivero Vaccari, J.P.; Keane, R.W. Human astrocytes express a novel NLRP2 inflammasome. Glia 2013, 61, 1113–1121. [Google Scholar] [CrossRef]
  125. de Rivero Vaccari, J.P.; Dietrich, W.D.; Keane, R.W. Activation and regulation of cellular inflammasomes: Gaps in our knowledge for central nervous system injury. J. Cereb. Blood Flow Metab. 2014, 34, 369–375. [Google Scholar] [CrossRef] [Green Version]
  126. Nyúl-Tóth, Á.; Kozma, M.; Nagyőszi, P.; Nagy, K.; Fazakas, C.; Haskó, J.; Molnár, K.; Farkas, A.E.; Végh, A.G.; Váró, G.; et al. Expression of pattern recognition receptors and activation of the non-canonical inflammasome pathway in brain pericytes. Brain Behav. Immun. 2017, 64, 220–231. [Google Scholar] [CrossRef] [PubMed]
  127. Johann, S.; Heitzer, M.; Kanagaratnam, M.; Goswami, A.; Rizo, T.; Weis, J.; Troost, D.; Beyer, C. NLRP3 inflammasome is expressed by astrocytes in the SOD1 mouse model of ALS and in human sporadic ALS patients. Glia 2015, 63, 2260–2273. [Google Scholar] [CrossRef]
  128. Ebrahimi, T.; Rust, M.; Kaiser, S.N.; Slowik, A.; Beyer, C.; Koczulla, A.R.; Schulz, J.B.; Habib, P.; Bach, J.P. α1-antitrypsin mitigates NLRP3-inflammasome activation in amyloid β1–42-stimulated murine astrocytes. J. Neuroinflamm. 2018, 15, 282. [Google Scholar] [CrossRef]
  129. Sandhu, J.K.; Kulka, M. Decoding Mast Cell-Microglia Communication in Neurodegenerative Diseases. Int. J. Mol. Sci. 2021, 22, 1093. [Google Scholar] [CrossRef]
  130. Komleva, Y.K.; Lopatina, O.L.; Gorina, Y.V.; Chernykh, A.I.; Trufanova, L.V.; Vais, E.F.; Kharitonova, E.V.; Zhukov, E.L.; Vahtina, L.Y.; Medvedeva, N.N.; et al. Expression of NLRP3 Inflammasomes in Neurogenic Niche Contributes to the Effect of Spatial Learning in Physiological Conditions but Not in Alzheimer′s Type Neurodegeneration. Cell. Mol. Neurobiol. 2021, 42, 1355–1371. [Google Scholar] [CrossRef]
  131. Saresella, M.; La Rosa, F.; Piancone, F.; Zoppis, M.; Marventano, I.; Calabrese, E.; Rainone, V.; Nemni, R.; Mancuso, R.; Clerici, M. The NLRP3 and NLRP1 inflammasomes are activated in Alzheimers disease. Mol. Neurodegener. 2016, 11, 23. [Google Scholar] [CrossRef] [Green Version]
  132. Tang, H.; Harte, M. Investigating markers of the NLRP3 inflammasome pathway in Alzheimer′s disease: A human post-mortem study. Genes 2021, 12, 1753. [Google Scholar] [CrossRef]
  133. Dal Prà, I.; Armato, U.; Chiarini, A. Family C G-Protein-Coupled Receptors in Alzheimer′s Disease and Therapeutic Implications. Front. Pharmacol. 2019, 10, 1282. [Google Scholar] [CrossRef]
  134. Dal Prà, I.; Armato, U.; Chioffi, F.; Pacchiana, R.; Whitfield, J.F.; Chakravarthy, B.; Gui, L.; Chiarini, A. The Aβ peptides-activated calcium-sensing receptor stimulates the production and secretion of vascular endothelial growth factor-A by normoxic adult human cortical astrocytes. Neuromol. Med. 2014, 16, 645–657. [Google Scholar] [CrossRef]
  135. Dal Prà, I.; Chiarini, A.; Pacchiana, R.; Gardenal, E.; Chakravarthy, B.; Whitfield, J.F.; Armato, U. Calcium-sensing receptors of human Astrocyte-Neuron Teams: Amyloid-β-driven mediators and therapeutic targets of Alzheimer′s Disease. Curr. Neuropharmacol. 2014, 12, 353–364. [Google Scholar] [CrossRef] [Green Version]
  136. Dal Prà, I.; Armato, U.; Chiarini, A. Specific interactions of calcium-sensing receptors (CaSRs) with soluble amyloid-β peptides—A study using cultured normofunctioning adult human astrocytes. In Proceedings of the 2nd International Symposium on the Calcium-Sensing Receptor, San Diego, CA, USA, 3–4 March 2015; pp. 90–91. [Google Scholar]
  137. Hofer, A.M.; Brown, E.M. Extracellular calcium sensing and signaling. Nat. Rev. Mol. Cell Biol. 2003, 4, 530–538. [Google Scholar] [CrossRef]
  138. Gardenal, E.; Chiarini, A.; Armato, U.; Dal Prà, I.; Verkhratsky, A.; Rodríguez, J.J. Increased calcium-sensing receptor immunoreactivity in the hippocampus of a triple transgenic mouse model of Alzheimer′s Disease. Front. Neurosci. 2017, 11, 81. [Google Scholar] [CrossRef] [Green Version]
  139. Gutiérrez-López, T.Y.; Orduña-Castillo, L.B.; Hernández-Vásquez, M.N.; Vázquez-Prado, J.; Reyes-Cruz, G. Calcium sensing receptor activates the NLRP3 inflammasome via a chaperone-assisted degradative pathway involving Hsp70 and LC3-II. Biochem. Biophys. Res. Commun. 2018, 505, 1121–1127. [Google Scholar] [CrossRef]
  140. Sokolowska, M.; Chen, L.Y.; Liu, Y.; Martinez-Anton, A.; Qi, H.Y.; Logun, C.; Alsaaty, S.; Park, Y.H.; Kastner, D.L.; Chae, J.J.; et al. Prostaglandin E2 Inhibits NLRP3 Inflammasome Activation through EP4 Receptor and Intracellular Cyclic AMP in Human Macrophages. J. Immunol. 2015, 194, 5472–5487. [Google Scholar] [CrossRef] [Green Version]
  141. Armato, U.; Chiarini, A.; Chakravarthy, B.; Chioffi, F.; Pacchiana, R.; Colarusso, E.; Whitfield, J.F.; Dal Prà, I. Calcium-sensing receptor antagonist (calcilytic) NPS 2143 specifically blocks the increased secretion of endogenous Aβ42 prompted by exogenous fibrillary or soluble Aβ25-35 in human cortical astrocytes and neurons-therapeutic relevance to Alzheimer′s disease. Biochim. Biophys. Acta 2013, 1832, 1634–1652. [Google Scholar] [CrossRef]
  142. Pi, M.; Faber, P.; Ekema, G.; Jackson, P.D.; Ting, A.; Wang, N.; Fontilla-Poole, M.; Mays, R.W.; Brunden, K.R.; Harrington, J.J.; et al. Identification of a novel extracellular cation-sensing G-protein-coupled receptor. J. Biol. Chem. 2005, 280, 40201–40209. [Google Scholar] [CrossRef] [Green Version]
  143. Pi, M.; Parrill, A.L.; Quarles, L.D. GPRC6A mediates the non-genomic effects of steroids. J. Biol. Chem. 2010, 285, 39953–39964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Pi, M.; Wu, Y.; Quarles, L.D. GPRC6A mediates responses to osteocalcin in β-cells in vitro and pancreas in vivo. J. Bone Miner. Res. 2011, 26, 1680–1683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Pi, M.; Quarles, L.D. GPRC6A regulates prostate cancer progression. Prostate 2011, 72, 399–409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Singh, P.; Dutta, S.R.; Song, C.Y.; Oh, S.; Gonzalez, F.J.; Malik, K.U. Brain Testosterone-CYP1B1 (Cytochrome P450 1B1) Generated Metabolite 6β-Hydroxytestosterone Promotes Neurogenic Hypertension and Inflammation. Hypertension 2020, 76, 1006–1018. [Google Scholar] [CrossRef] [PubMed]
  147. Bai, N.; Zhang, Q.; Zhang, W.; Liu, B.; Yang, F.; Brann, D.; Wang, R. G-protein-coupled estrogen receptor activation upregulates interleukin-1 receptor antagonist in the hippocampus after global cerebral ischemia: Implications for neuronal self-defense. J. Neuroinflamm. 2020, 17, 45. [Google Scholar] [CrossRef]
  148. Py, B.F.; Kim, M.S.; Vakifahmetoglu-Norberg, H.; Yuan, J. Deubiquitination of NLRP3 by BRCC3 critically regulates inflammasome activity. Mol. Cell. 2013, 49, 331–338. [Google Scholar] [CrossRef] [Green Version]
  149. Yang, J.; Wise, L.; Fukuchi, K.I. TLR4 Cross-Talk with NLRP3 Inflammasome and Complement Signaling Pathways in Alzheimer′s Disease. Front. Immunol. 2020, 11, 724. [Google Scholar] [CrossRef]
  150. McKee, C.M.; Coll, R.C. NLRP3 inflammasome priming: A riddle wrapped in a mystery inside an enigma. J. Leukoc. Biol. 2020, 108, 937–952. [Google Scholar] [CrossRef]
  151. Chen, M.-Y.; Ye, X.J.; He, X.H.; Ouyang, D.Y. The Signaling Pathways Regulating NLRP3 Inflammasome Activation. Inflammation 2021, 44, 1229–1245. [Google Scholar] [CrossRef]
  152. Dierckx, T.; Haidar, M.; Grajchen, E.; Wouters, E.; Vanherle, S.; Loix, M.; Boeykens, A.; Bylemans, D.; Hardonnière, K.; Kerdine-Römer, S.; et al. Phloretin suppresses neuroinflammation by autophagy-mediated Nrf2 activation in macrophages. J. Neuroinflamm. 2021, 18, 148. [Google Scholar] [CrossRef]
  153. Katsnelson, M.A.; Rucker, L.G.; Russo, H.M.; Dubyak, G.R. K+ efflux agonists induce NLRP3 inflammasome activation independently of Ca2+ signaling. J. Immunol. 2015, 194, 3937–3952. [Google Scholar] [CrossRef] [Green Version]
  154. Elliott, E.I.; Sutterwala, F.S. Initiation and perpetuation of NLRP3 inflammasome activation and assembly. Immunol. Rev. 2015, 265, 35–52. [Google Scholar] [CrossRef] [Green Version]
  155. Zhou, Y.; Tong, Z.; Jiang, S.; Zheng, W.; Zhao, J.; Zhou, X. The Roles of Endoplasmic Reticulum in NLRP3 Inflammasome Activation. Cells 2020, 9, 1219. [Google Scholar] [CrossRef]
  156. Jäger, E.; Murthy, S.; Schmidt, C.; Hahn, M.; Strobel, S.; Peters, A.; Stäubert, C.; Sungur, P.; Venus, T.; Geisler, M.; et al. Calcium-sensing receptor-mediated NLRP3 inflammasome response to calciprotein particles drives inflammation in rheumatoid arthritis. Nat. Commun. 2020, 11, 4243. [Google Scholar] [CrossRef]
  157. Murakami, T.; Ockinger, J.; Yu, J.; Byles, V.; McColl, A.; Hofer, A.M.; Horng, T. Critical role for calcium mobilization in activation of the NLRP3 inflammasome. Proc. Natl. Acad. Sci. USA 2012, 109, 11282–11287. [Google Scholar] [CrossRef] [Green Version]
  158. Rossol, M.; Pierer, M.; Raulien, N.; Quandt, D.; Meusch, U.; Rothe, K.; Schubert, K.; Schöneberg, T.; Schaefer, M.; Krügel, U.; et al. Extracellular Ca2+ is a danger signal activating the NLRP3 inflammasome through G protein-coupled calcium sensing receptors. Nat. Commun. 2012, 3, 1329. [Google Scholar] [CrossRef] [Green Version]
  159. Kim, K.; Kim, H.J.; Binas, B.; Kang, J.H.; Chung, I.Y. Inflammatory mediators ATP and S100A12 activate the NLRP3 inflammasome to induce MUC5AC production in airway epithelial cells. Biochem. Biophys. Res. Commun. 2018, 503, 657–664. [Google Scholar] [CrossRef]
  160. 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]
  161. Lim, J.C.; Lu, W.; Beckel, J.M.; Mitchell, C.H. Neuronal Release of Cytokine IL-3 Triggered by Mechanosensitive Autostimulation of the P2X7 Receptor Is Neuroprotective. Front. Cell. Neurosci. 2016, 10, 270. [Google Scholar] [CrossRef] [Green Version]
  162. Lu, W.; Albalawi, F.; Beckel, J.M.; Lim, J.C.; Laties, A.M.; Mitchell, C.H. The P2X7 receptor links mechanical strain to cytokine IL-6 up-regulation and release in neurons and astrocytes. J. Neurochem. 2017, 141, 436–448. [Google Scholar] [CrossRef] [Green Version]
  163. Campagno, K.E.; Mitchell, C.H. The P2X7Receptor in Microglial Cells Modulates the Endolysosomal Axis, Autophagy, and Phagocytosis. Front. Cell. Neurosci. 2021, 15, 645244. [Google Scholar] [CrossRef] [PubMed]
  164. Shieh, C.H.; 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]
  165. Cieślak, M.; Wojtczak, A. Role of purinergic receptors in the Alzheimer′s disease. Purinergic Signal. 2018, 14, 331–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Erb, L.; Woods, L.T.; Khalafalla, M.G.; Weisman, G.A. Purinergic signaling in Alzheimer′s disease. Brain Res. Bull. 2019, 151, 25–37. [Google Scholar] [CrossRef]
  167. Duez, H.; Pourcet, B. Nuclear Receptors in the Control of the NLRP3 Inflammasome Pathway. Front. Endocrinol. 2021, 12, 630536. [Google Scholar] [CrossRef]
  168. Swanson, K.V.; Deng, M.; Ting, J.P. The NLRP3 inflammasome: Molecular activation and regulation to therapeutics. Nat. Rev. Immunol. 2019, 19, 477–489. [Google Scholar] [CrossRef]
  169. Liang, Z.; Damianou, A.; Di Daniel, E.; Kessler, B.M. Inflammasome activation controlled by the interplay between post-translational modifications: Emerging drug target opportunities. Cell Commun. Signal. 2021, 19, 23. [Google Scholar] [CrossRef]
  170. Weber, A.N.R. Targeting the NLRP3 Inflammasome via BTK. Front. Cell Dev. Biol. 2021, 9, 630479. [Google Scholar] [CrossRef]
  171. Bezbradica, J.S.; Coll, R.C.; Schroder, K. Sterile signals generate weaker and delayed macrophage NLRP3 inflammasome responses relative to microbial signals. Cell. Mol. Immunol. 2017, 14, 118–126. [Google Scholar] [CrossRef] [Green Version]
  172. Healy, L.M.; Yaqubi, M.; Ludwin, S.; Antel, J.P. Species differences in immune-mediated CNS tissue injury and repair: A (neuro)inflammatory topic. Glia 2020, 68, 811–829. [Google Scholar] [CrossRef]
  173. Zhang, C.J.; Jiang, M.; Zhou, H.; Liu, W.; Wang, C.; Kang, Z.; Han, B.; Zhang, Q.; Chen, X.; Xiao, J.; et al. TLR-stimulated IRAKM activates caspase-8 inflammasome in microglia and promotes neuroinflammation. J. Clin. Investing. 2018, 128, 5399–5412. [Google Scholar] [CrossRef]
  174. Kayagaki, N.; Warming, S.; Lamkanfi, M.; Vande Walle, L.; Louie, S.; Dong, J.; Newton, K.; Qu, Y.; Liu, J.; Heldens, S.; et al. Non-canonical inflammasome activation targets caspase-11. Nature 2011, 479, 117–121. [Google Scholar] [CrossRef] [PubMed]
  175. Elizagaray, M.L.; Gomes, M.T.R.; Guimaraes, E.S.; Rumbo, M.; Hozbor, D.F.; Oliveira, S.C.; Moreno, G. Canonical and Non-canonical Inflammasome Activation by Outer Membrane Vesicles Derived from Bordetella pertussis. Front. Immunol. 2020, 11, 1879. [Google Scholar] [CrossRef]
  176. Matikainen, S.; Nyman, T.A.; Cypryk, W. Function and Regulation of Noncanonical Caspase-4/5/11 Inflammasome. J. Immunol. 2020, 204, 3063–3069. [Google Scholar] [CrossRef]
  177. Yi, Y.S. Caspase-11 Noncanonical Inflammasome: A Novel Key Player in Murine Models of Neuroinflammation and Multiple Sclerosis. Neuroimmunomodulation 2021, 28, 195–203. [Google Scholar] [CrossRef]
  178. Zhang, D.; Qian, J.; Zhang, P.; Li, H.; Shen, H.; Li, X.; Chen, G. Gasdermin D serves as a key executioner of pyroptosis in experimental cerebral ischemia and reperfusion model both in vivo and in vitro. J. Neurosci. Res. 2019, 97, 645–660. [Google Scholar] [CrossRef]
  179. Wang, K.; Sun, Z.; Ru, J.; Wang, S.; Huang, L.; Ruan, L.; Lin, X.; Jin, K.; Zhuge, Q.; Yang, S. Ablation of GSDMD Improves Outcome of Ischemic Stroke Through Blocking Canonical and Non-canonical Inflammasomes Dependent Pyroptosis in Microglia. Front. Neurol. 2020, 11, 577927. [Google Scholar] [CrossRef]
  180. Carpenter, S.; Aiello, D.; Atianand, M.K.; Ricci, E.P.; Gandhi, P.; Hall, L.L.; Byron, M.; Monks, B.; Henry-Bezy, M.; Lawrence, J.B.; et al. A long noncoding RNA mediates both activation and repression of immune response genes. Science 2013, 341, 789–792. [Google Scholar] [CrossRef] [Green Version]
  181. Heward, J.A.; Lindsay, M.A. Long non-coding RNAs in the regulation of the immune response. Trends Immunol. 2014, 35, 408–419. [Google Scholar] [CrossRef] [Green Version]
  182. Bartel, D.P. Metazoan MicroRNAs. Cell 2018, 173, 20–51. [Google Scholar] [CrossRef] [Green Version]
  183. Kiesel, P.; Gibson, T.J.; Ciesielczyk, B.; Bodemer, M.; Kaup, F.J.; Bodemer, W.; Zischler, H.; Zerr, I. Transcription of Alu DNA elements in blood cells of sporadic Creutzfeldt-Jakob disease (sCJD). Prion 2010, 4, 87–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Polesskaya, O.; Kananykhina, E.; Roy-Engel, A.M.; Nazarenko, O.; Kulemzina, I.; Baranova, A.; Vassetsky, Y.; Myakishev-Rempel, M. The role of Alu-derived RNAs in Alzheimer′s and other neurodegenerative conditions. Med. Hypotheses 2018, 115, 29–34. [Google Scholar] [CrossRef] [PubMed]
  185. Cheng, Y.; Saville, L.; Gollen, B.; Isaac, C.; Belay, A.; Mehla, J.; Patel, K.; Thakor, N.; Mohajerani, M.H.; Zovoilis, A. Increased processing of SINE B2 ncRNAs unveils a novel type of transcriptome deregulation in amyloid beta neuropathology. eLife 2020, 9, e61265. [Google Scholar] [CrossRef]
  186. Cheng, Y.; Saville, L.; Gollen, B.; Veronesi, A.A.; Mohajerani, M.; Joseph, J.T.; Zovoilis, A. Increased Alu RNA processing in Alzheimer brains is linked to gene expression changes. EMBO Rep. 2021, 22, e52255. [Google Scholar] [CrossRef]
  187. Zhao, Y.; Chen, Y.; Wang, Z.; Xu, C.; Qiao, S.; Liu, T.; Qi, K.; Tong, D.; Li, C. Bone Marrow Mesenchymal Stem Cell Exosome Attenuates Inflammasome-Related Pyroptosis via Delivering circ_003564 to Improve the Recovery of Spinal Cord Injury. Mol. Neurobiol. 2022, 59, 6771–6789. [Google Scholar] [CrossRef]
  188. Xue, Z.; Zhang, Z.; Liu, H.; Li, W.; Guo, X.; Zhang, Z.; Liu, Y.; Jia, L.; Li, Y.; Ren, Y.; et al. lincRNA-Cox2 regulates NLRP3 inflammasome and autophagy mediated neuroinflammation. Cell Death Differ. 2019, 26, 130–145. [Google Scholar] [CrossRef] [Green Version]
  189. Meng, J.; Ding, T.; Chen, Y.; Long, T.; Xu, Q.; Lian, W.; Liu, W. LncRNA-Meg3 promotes Nlrp3-mediated microglial inflammation by targeting miR-7a-5p. Int. Immunopharmacol. 2021, 90, 107141. [Google Scholar] [CrossRef]
  190. Docrat, T.F.; Nagiah, S.; Chuturgoon, A.A. Metformin protects against neuroinflammation through integrated mechanisms of miR-141 and the NF-ĸB-mediated inflammasome pathway in a diabetic mouse model. Eur. J. Pharmacol. 2021, 903, 174146. [Google Scholar] [CrossRef]
  191. Cunha, C.; Gomes, C.; Vaz, A.R.; Brites, D. Exploring New Inflammatory Biomarkers and Pathways during LPS-Induced M1 Polarization. Mediat. Inflamm. 2016, 2016, 6986175. [Google Scholar] [CrossRef] [Green Version]
  192. Si, L.; Wang, H.; Wang, L. Suppression of miR-193a alleviates neuroinflammation and improves neurological function recovery after traumatic brain injury (TBI) in mice. Biochem. Biophys. Res. Commun. 2020, 523, 527–534. [Google Scholar] [CrossRef]
  193. Cao, Y.; Tan, X.; Lu, Q.; Huang, K.; Tang, X.; He, Z. miR-590-3 and SP1 Promote Neuronal Apoptosis in Patients with Alzheimer′s Disease via AMPK Signaling Pathway. Contrast Media Mol. Imaging 2021, 2021, 6010362. [Google Scholar] [CrossRef] [PubMed]
  194. Zhang, H.; Tao, J.; Zhang, S.; Lv, X. LncRNA MEG3 Reduces Hippocampal Neuron Apoptosis via the PI3K/AKT/mTOR Pathway in a Rat Model of Temporal Lobe Epilepsy. Neuropsychiatr. Dis. Treat. 2020, 16, 2519–2528. [Google Scholar] [CrossRef]
  195. Fan, Z.; Lu, M.; Qiao, C.; Zhou, Y.; Ding, J.H.; Hu, G. MicroRNA-7 Enhances Subventricular Zone Neurogenesis by Inhibiting NLRP3/Caspase-1 Axis in Adult Neural Stem Cells. Mol. Neurobiol. 2016, 53, 7057–7069. [Google Scholar] [CrossRef]
  196. Zhou, Y.; Lu, M.; Du, R.H.; Qiao, C.; Jiang, C.Y.; Zhang, K.Z.; Ding, J.H.; Hu, G. MicroRNA-7 targets Nod-like receptor protein 3 inflammasome to modulate neuroinflammation in the pathogenesis of Parkinson′s disease. Mol. Neurodegener. 2016, 11, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Cui, G.H.; Wu, J.; Mou, F.F.; Xie, W.H.; Wang, F.B.; Wang, Q.L.; Fang, J.; Xu, Y.W.; Dong, Y.R.; Liu, J.R.; et al. Exosomes derived from hypoxia-preconditioned mesenchymal stromal cells ameliorate cognitive decline by rescuing synaptic dysfunction and regulating inflammatory responses in APP/PS1 mice. FASEB J. 2018, 32, 654–668. [Google Scholar] [CrossRef] [Green Version]
  198. Han, C.; Guo, L.; Yang, Y.; Guan, Q.; Shen, H.; Sheng, Y.; Jiao, Q. Mechanism of microRNA-22 in regulating neuroinflammation in Alzheimer′s disease. Brain Behav. 2020, 10, e01627. [Google Scholar] [CrossRef] [Green Version]
  199. Zhai, L.; Shen, H.; Sheng, Y.; Guan, Q. ADMSC Exo-MicroRNA-22 improve neurological function and neuroinflammation in mice with Alzheimer′s disease. J. Cell. Mol. Med. 2021, 25, 7513–7523, Erratum in J. Cell. Mol. Med. 2021, 25, 11037–11038. [Google Scholar] [CrossRef]
  200. Hu, L.T.; Wang, B.Y.; Fan, Y.H.; He, Z.Y.; Zheng, W.X. Exosomal miR-23b from bone marrow mesenchymal stem cells alleviates oxidative stress and pyroptosis after intracerebral hemorrhage. Neural Regen. Res. 2023, 18, 560–567. [Google Scholar] [CrossRef]
  201. Cao, Y.; Tan, X.; Lu, Q.; Huang, K.; Tang, X.; He, Z. MiR-29c-3p May Promote the Progression of Alzheimer′s Disease through BACE1. J. Healthc. Eng. 2021, 2021, 2031407. [Google Scholar] [CrossRef]
  202. Sha, S.; Shen, X.; Cao, Y.; Qu, L. Mesenchymal stem cells-derived extracellular vesicles ameliorate Alzheimer′s disease in rat models via the microRNA-29c-3p/BACE1 axis and the Wnt/β-catenin pathway. Aging 2021, 13, 15285–15306. [Google Scholar] [CrossRef]
  203. Hu, L.; Zhang, H.; Wang, B.; Ao, Q.; He, Z. MicroRNA-152 attenuates neuroinflammation in intracerebral hemorrhage by inhibiting thioredoxin interacting protein (TXNIP)-mediated NLRP3 inflammasome activation. Int. Immunopharmacol. 2020, 80, 106141. [Google Scholar] [CrossRef] [PubMed]
  204. Li, Q.; Wang, Z.; Xing, H.; Wang, Y.; Guo, Y. Exosomes derived from miR-188-3p-modified adipose-derived mesenchymal stem cells protect Parkinson′s disease. Mol. Ther. Nucleic Acids 2021, 23, 1334–1344. [Google Scholar] [CrossRef] [PubMed]
  205. Wan, S.Y.; Li, G.S.; Tu, C.; Chen, W.L.; Wang, X.W.; Wang, Y.N.; Peng, L.B.; Tan, F. MicroNAR-194-5p hinders the activation of NLRP3 inflammasomes and alleviates neuroinflammation during intracerebral hemorrhage by blocking the interaction between TRAF6 and NLRP3. Brain Res. 2021, 1752, 147228. [Google Scholar] [CrossRef] [PubMed]
  206. Mancuso, R.; Agostini, S.; Hernis, A.; Zanzottera, M.; Bianchi, A.; Clerici, M. Circulatory miR-223-3p Discriminates Between Parkinson′s and Alzheimer′s Patients. Sci. Rep. 2019, 9, 9393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Chen, Z.; Hu, Y.; Lu, R.; Ge, M.; Zhang, L. MicroRNA-374a-5p inhibits neuroinflammation in neonatal hypoxic-ischemic encephalopathy via regulating NLRP3 inflammasome targeted Smad6. Life Sci. 2020, 252, 117664. [Google Scholar] [CrossRef]
  208. Kaur, S.; Verma, H.; Dhiman, M.; Tell, G.; Gigli, G.L.; Janes, F.; Mantha, A.K. Brain Exosomes: Friend or Foe in Alzheimer′s Disease? Mol. Neurobiol. 2021, 58, 6610–6624. [Google Scholar] [CrossRef]
  209. Hu, Z.; Yuan, Y.; Zhang, X.; Lu, Y.; Dong, N.; Jiang, X.; Xu, J.; Zheng, D. Human Umbilical Cord Mesenchymal Stem Cell-Derived Exosomes Attenuate Oxygen-Glucose Deprivation/Reperfusion-Induced Microglial Pyroptosis by Promoting FOXO3a-Dependent Mitophagy. Oxid. Med. Cell. Longev. 2021, 2021, 6219715. [Google Scholar] [CrossRef]
  210. Liu, X.; Zhang, M.; Liu, H.; Zhu, R.; He, H.; Zhou, Y.; Zhang, Y.; Li, C.; Liang, D.; Zeng, Q.; et al. Bone marrow mesenchymal stem cell-derived exosomes attenuate cerebral ischemia-reperfusion injury-induced neuroinflammation and pyroptosis by modulating microglia M1/M2 phenotypes. Exp. Neurol. 2021, 341, 113700. [Google Scholar] [CrossRef]
  211. Cui, G.H.; Guo, H.D.; Li, H.; Zhai, Y.; Gong, Z.B.; Wu, J.; Liu, J.S.; Dong, Y.R.; Hou, S.X.; Liu, J.R. RVG-modified exosomes derived from mesenchymal stem cells rescue memory deficits by regulating inflammatory responses in a mouse model of Alzheimer′s disease. Immun. Ageing 2019, 16, 10. [Google Scholar] [CrossRef] [Green Version]
  212. Chiarini, A.; Armato, U.; Gardenal, E.; Gui, L.; Dal Prà, I. Amyloid β-exposed human astrocytes overproduce phospho-Tau and overrelease it within exosomes, effects suppressed by calcilytic NPS 2143. Further implications for Alzheimer′s therapy. Front. Neurosci. 2017, 11, 217. [Google Scholar] [CrossRef] [Green Version]
  213. Sardar Sinha, M.; Ansell-Schultz, A.; Civitelli, L.; Hildesjö, C.; Larsson, M.; Lannfelt, L.; Ingelsson, M.; Hallbeck, M. Alzheimer′s disease pathology propagation by exosomes containing toxic amyloid-beta oligomers. Acta Neuropathol. 2018, 136, 41–56. [Google Scholar] [CrossRef] [Green Version]
  214. Zhang, Q.; Sun, Y.; He, Z.; Xu, Y.; Li, X.; Ding, J.; Lu, M.; Hu, G. Kynurenine regulates NLRP2 inflammasome in astrocytes and its implications in depression. Brain Behav. Immun. 2020, 88, 471–481. [Google Scholar] [CrossRef]
  215. Voet, S.; Mc Guire, C.; Hagemeyer, N.; Martens, A.; Schroeder, A.; Wieghofer, P.; Daems, C.; Staszewski, O.; Vande Walle, L.; Jordao, M.J.C.; et al. A20 critically controls microglia activation and inhibits inflammasome-dependent neuroinflammation. Nat. Commun. 2018, 9, 2036. [Google Scholar] [CrossRef] [Green Version]
  216. Gaikwad, S.; Patel, D.; Agrawal-Rajput, R. CD40 Negatively Regulates ATP-TLR4-Activated Inflammasome in Microglia. Cell. Mol. Neurobiol. 2017, 37, 351–359. [Google Scholar] [CrossRef]
  217. Ma, S.; Wang, Y.; Zhou, X.; Li, Z.; Zhang, Z.; Wang, Y.; Huang, T.; Zhang, Y.; Shi, J.; Guan, F. MG53 Protects hUC-MSCs against Inflammatory Damage and Synergistically Enhances Their Efficacy in Neuroinflammation Injured Brain through Inhibiting NLRP3/Caspase-1/IL-1β Axis. ACS Chem. Neurosci. 2020, 11, 2590–2601. [Google Scholar] [CrossRef]
  218. Xiao, T.; Wan, J.; Qu, H.; Li, Y. Tripartite-motif protein 21 knockdown extenuates LPS-triggered neurotoxicity by inhibiting microglial M1 polarization via suppressing NF-κB-mediated NLRP3 inflammasome activation. Arch. Biochem. Biophys. 2021, 706, 108918. [Google Scholar] [CrossRef]
  219. Gal-Ben-Ari, S.; Barrera, I.; Ehrlich, M.; Rosenblum, K. PKR: A Kinase to Remember. Front. Mol. Neurosci. 2019, 11, 480. [Google Scholar] [CrossRef] [Green Version]
  220. Lu, B.; Nakamura, T.; Inouye, K.; Li, J.; Tang, Y.; Lundbäck, P.; Valdes-Ferrer, S.I.; Olofsson, P.S.; Kalb, T.; Roth, J.; et al. Novel role of PKR in inflammasome activation and HMGB1 release. Nature 2012, 488, 670–674. [Google Scholar] [CrossRef] [Green Version]
  221. He, Y.; Franchi, L.; Núñez, G. The protein kinase PKR is critical for LPS-induced iNOS production but dispensable for inflammasome activation in macrophages. Eur. J. Immunol. 2013, 43, 1147–1152. [Google Scholar] [CrossRef] [Green Version]
  222. Dempsey, C.; Rubio Araiz, A.; Bryson, K.J.; Finucane, O.; Larkin, C.; Mills, E.L.; Robertson, A.; Cooper, M.A.; O′Neill, L.; Lynch, M.A. Inhibiting the NLRP3 inflammasome with MCC950 promotes non-phlogistic clearance of amyloid-β and cognitive function in APP/PS1 mice. Brain Behav. Immun. 2017, 61, 306–316. [Google Scholar] [CrossRef] [Green Version]
  223. Heitzer, M.; Kaiser, S.; Kanagaratnam, M.; Zendedel, A.; Hartmann, P.; Beyer, C.; Johann, S. Administration of 17beta-Estradiol Improves Motoneuron Survival and Down-regulates Inflammasome Activation in Male SOD1 (G93A) ALS Mice. Mol. Neurobiol. 2017, 54, 8429–8443. [Google Scholar] [CrossRef] [PubMed]
  224. Aryanpour, R.; Zibara, K.; Pasbakhsh, P.; Jame′ei, S.B.; Namjoo, Z.; Ghanbari, A.; Mahmoudi, R.; Amani, S.; Kashani, I.R. 17β-Estradiol Reduces Demyelination in Cuprizone-fed Mice by Promoting M2 Microglia Polarity and Regulating NLRP3 Inflammasome. Neuroscience 2021, 463, 116–127. [Google Scholar] [CrossRef] [PubMed]
  225. Thakkar, R.; Wang, R.; Wang, J.; Vadlamudi, R.K.; Brann, D.W. 17β-Estradiol Regulates Microglia Activation and Polarization in the Hippocampus Following Global Cerebral Ischemia. Oxid. Med. Cell. Longev. 2018, 2018, 4248526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Jiang, W.; Huang, Y.; He, F.; Liu, J.; Li, M.; Sun, T.; Ren, W.; Hou, J.; Zhu, L. Dopamine D1 Receptor Agonist A-68930 Inhibits NLRP3 Inflammasome Activation, Controls Inflammation, and Alleviates Histopathology in a Rat Model of Spinal Cord Injury. Spine (Phila Pa 1976) 2016, 41, E330–E334. [Google Scholar] [CrossRef]
  227. Wang, S.; Yao, Q.; Wan, Y.; Wang, J.; Huang, C.; Li, D.; Yang, B. Adiponectin reduces brain injury after intracerebral hemorrhage by reducing NLRP3 inflammasome expression. Int. J. Neurosci. 2020, 130, 301–308. [Google Scholar] [CrossRef]
  228. Li, J.; Wu, D.M.; Yu, Y.; Deng, S.H.; Liu, T.; Zhang, T.; He, M.; Zhao, Y.Y.; Xu, Y. Amifostine ameliorates induction of experimental autoimmune encephalomyelitis: Effect on reactive oxygen species/NLRP3 pathway. Int. Immunopharmacol. 2020, 88, 106998. [Google Scholar] [CrossRef]
  229. Li, B.X.; Dai, X.; Xu, X.R.; Adili, R.; Neves, M.A.D.; Lei, X.; Shen, C.; Zhu, G.; Wang, Y.; Zhou, H.; et al. In vitro assessment and phase I randomized clinical trial of anfibatide a snake venom derived anti-thrombotic agent targeting human platelet GPIbα. Sci. Rep. 2021, 11, 11663. [Google Scholar] [CrossRef]
  230. Li, R.; Si, M.; Jia, H.Y.; Ma, Z.; Li, X.W.; Li, X.Y.; Dai, X.R.; Gong, P.; Luo, S.Y. Anfibatide alleviates inflammation and apoptosis via inhibiting NF-kappaB/NLRP3 axis in ischemic stroke. Eur. J. Pharmacol. 2022, 926, 175032. [Google Scholar] [CrossRef]
  231. Liu, P.; Gao, Q.; Guan, L.; Hu, Y.; Jiang, J.; Gao, T.; Sheng, W.; Xue, X.; Qiao, H.; Li, T. Atorvastatin attenuates surgery-induced BBB disruption and cognitive impairment partly by suppressing NF-κB pathway and NLRP3 inflammasome activation in aged mice. Acta Biochim. Biophys. Sin. 2021, 53, 528–537. [Google Scholar] [CrossRef]
  232. Jiang, W.; Li, M.; He, F.; Zhou, S.; Zhu, L. Targeting the NLRP3 inflammasome to attenuate spinal cord injury in mice. J. Neuroinflamm. 2017, 14, 207. [Google Scholar] [CrossRef] [Green Version]
  233. Yang, T.; Zhang, L.; Shang, Y.; Zhu, Z.; Jin, S.; Guo, Z.; Wang, X. Concurrent suppression of Aβ aggregation and NLRP3 inflammasome activation for treating Alzheimer′s disease. Chem. Sci. 2022, 13, 2971–2980. [Google Scholar] [CrossRef]
  234. Wang, H.Q.; Song, K.Y.; Feng, J.Z.; Huang, S.Y.; Guo, X.M.; Zhang, L.; Zhang, G.; Huo, Y.C.; Zhang, R.R.; Ma, Y.; et al. Caffeine Inhibits Activation of the NLRP3 Inflammasome via Autophagy to Attenuate Microglia-Mediated Neuroinflammation in Experimental Autoimmune Encephalomyelitis. J. Mol. Neurosci. 2022, 72, 97–112. [Google Scholar] [CrossRef]
  235. de Oliveira, L.R.C.; Mimura, L.A.N.; Fraga-Silva, T.F.C.; Ishikawa, L.L.W.; Fernandes, A.A.H.; Zorzella-Pezavento, S.F.G.; Sartori, A. Calcitriol Prevents Neuroinflammation and Reduces Blood-Brain Barrier Disruption and Local Macrophage/Microglia Activation. Front. Pharmacol. 2020, 11, 161. [Google Scholar] [CrossRef]
  236. Wang, Y.; Guan, X.; Chen, X.; Cai, Y.; Ma, Y.; Ma, J.; Zhang, Q.; Dai, L.; Fan, X.; Bai, Y. Choline Supplementation Ameliorates Behavioral Deficits and Alzheimer′s Disease-Like Pathology in Transgenic APP/PS1 Mice. Mol. Nutr. Food Res. 2019, 63, e1801407. [Google Scholar] [CrossRef]
  237. Lonnemann, N.; Hosseini, S.; Marchetti, C.; Skouras, D.B.; Stefanoni, D.; D′Alessandro, A.; Dinarello, C.A.; Korte, M. The NLRP3 inflammasome inhibitor OLT1177 rescues cognitive impairment in a mouse model of Alzheimer′s disease. Proc. Natl. Acad. Sci. USA 2020, 117, 32145–32154. [Google Scholar] [CrossRef]
  238. Bao, Y.; Zhu, Y.; He, G.; Ni, H.; Liu, C.; Ma, L.; Zhang, L.; Shi, D. Dexmedetomidine Attenuates Neuroinflammation in LPS-Stimulated BV2 Microglia Cells Through Upregulation Of miR-340. Drug Des. Devel. Ther. 2019, 13, 3465–3475. [Google Scholar] [CrossRef] [Green Version]
  239. Feng, J.; Wang, J.X.; Du, Y.H.; Liu, Y.; Zhang, W.; Chen, J.F.; Liu, Y.J.; Zheng, M.; Wang, K.J.; He, G.Q. Dihydromyricetin inhibits microglial activation and neuroinflammation by suppressing NLRP3 inflammasome activation in APP/PS1 transgenic mice. CNS Neurosci. Ther. 2018, 24, 1207–1218. [Google Scholar] [CrossRef]
  240. Yan, Y.; Jiang, W.; Liu, L.; Wang, X.; Ding, C.; Tian, Z.; Zhou, R. Dopamine controls systemic inflammation through inhibition of NLRP3 inflammasome. Cell 2015, 160, 62–73. [Google Scholar] [CrossRef] [Green Version]
  241. Nizami, S.; Arunasalam, K.; Green, J.; Cook, J.; Lawrence, C.B.; Zarganes-Tzitzikas, T.; Davis, J.B.; Di Daniel, E.; Brough, D. Inhibition of the NLRP3 inflammasome by HSP90 inhibitors. Immunology 2021, 162, 84–91. [Google Scholar] [CrossRef]
  242. Gao, S.; Xu, T.; Guo, H.; Deng, Q.; Xun, C.; Liang, W.; Sheng, W. Ameliorative effects of echinacoside against spinal cord injury via inhibiting NLRP3 inflammasome signaling pathway. Life Sci. 2019, 237, 116978. [Google Scholar] [CrossRef]
  243. Kiasalari, Z.; Afshin-Majd, S.; Baluchnejadmojarad, T.; Azadi-Ahmadabadi, E.; Esmaeil-Jamaat, E.; Fahanik-Babaei, J.; Fakour, M.; Fereidouni, F.; Ghasemi-Tarie, R.; Jalalzade-Ogvar, S.; et al. Ellagic acid ameliorates neuroinflammation and demyelination in experimental autoimmune encephalomyelitis: Involvement of NLRP3 and pyroptosis. J. Chem. Neuroanat. 2021, 111, 101891. [Google Scholar] [CrossRef] [PubMed]
  244. Yang, X.; Sun, J.; Kim, T.J.; Kim, Y.J.; Ko, S.B.; Kim, C.K.; Jia, X.; Yoon, B.W. Pretreatment with low-dose fimasartan ameliorates NLRP3 inflammasome-mediated neuroinflammation and brain injury after intracerebral hemorrhage. Exp. Neurol. 2018, 310, 22–32. [Google Scholar] [CrossRef] [PubMed]
  245. Abu-Elfotuh, K.; Al-Najjar, A.H.; Mohammed, A.A.; Aboutaleb, A.S.; Badawi, G.A. Fluoxetine ameliorates Alzheimer′s disease progression and prevents the exacerbation of cardiovascular dysfunction of socially isolated depressed rats through activation of Nrf2/HO-1 and hindering TLR4/NLRP3 inflammasome signaling pathway. Int. Immunopharmacol. 2022, 104, 108488. [Google Scholar] [CrossRef] [PubMed]
  246. Liu, F.; Li, Z.; He, X.; Yu, H.; Feng, J. Ghrelin Attenuates Neuroinflammation and Demyelination in Experimental Autoimmune Encephalomyelitis Involving NLRP3 Inflammasome Signaling Pathway and Pyroptosis. Front. Pharmacol. 2019, 10, 1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Qu, J.; Tao, X.Y.; Teng, P.; Zhang, Y.; Guo, C.L.; Hu, L.; Qian, Y.N.; Jiang, C.Y.; Liu, W.T. Blocking ATP-sensitive potassium channel alleviates morphine tolerance by inhibiting HSP70-TLR4-NLRP3-mediated neuroinflammation. J. Neuroinflamm. 2017, 14, 228. [Google Scholar] [CrossRef] [Green Version]
  248. Hou, L.; Yang, J.; Li, S.; Huang, R.; Zhang, D.; Zhao, J.; Wang, Q. Glibenclamide attenuates 2.5-hexanedione-induced neurotoxicity in the spinal cord of rats through mitigation of NLRP3 inflammasome activation, neuroinflammation and oxidative stress. Toxicol. Lett. 2020, 331, 152–158. [Google Scholar] [CrossRef]
  249. Shao, B.Z.; Wei, W.; Ke, P.; Xu, Z.Q.; Zhou, J.X.; Liu, C. Activating cannabinoid receptor 2 alleviates pathogenesis of experimental autoimmune encephalomyelitis via activation of autophagy and inhibiting NLRP3 inflammasome. CNS Neurosci. Ther. 2014, 20, 1021–1028. [Google Scholar] [CrossRef]
  250. Karkhah, A.; Saadi, M.; Pourabdolhossein, F.; Saleki, K.; Nouri, H.R. Indomethacin attenuates neuroinflammation and memory impairment in an STZ-induced model of Alzheimer′s like disease. Immunopharmacol. Immunotoxicol. 2021, 43, 758–766. [Google Scholar] [CrossRef]
  251. Cooper, M.A. Inzomelid is a CNS penetrant anti-inflammatory drug that blocks NLRP3 inflammasome activation targeted to prevent Synuclein Pathology and Dopaminergic Degeneration in Parkinson′s disease. In Proceedings of the 7th International Conference on Parkinson′s & Movement Disorders, London, UK, 11–12 November 2019. [Google Scholar]
  252. Kuwar, R.; Rolfe, A.; Di, L.; Xu, H.; He, L.; Jiang, Y.; Zhang, S.; Sun, D. A novel small molecular NLRP3 inflammasome inhibitor alleviates neuroinflammatory response following traumatic brain injury. J. Neuroinflamm. 2019, 16, 81. [Google Scholar] [CrossRef] [Green Version]
  253. Lyu, D.; Wang, F.; Zhang, M.; Yang, W.; Huang, H.; Huang, Q.; Wu, C.; Qian, N.; Wang, M.; Zhang, H.; et al. Ketamine induces rapid antidepressant effects via the autophagy-NLRP3 inflammasome pathway. Psychopharmacology 2022, 239, 3201–3212. [Google Scholar] [CrossRef]
  254. Liu, S.; Wang, S.; Gu, R.; Che, N.; Wang, J.; Cheng, J.; Yuan, Z.; Cheng, Y.; Liao, Y. The XPO1 Inhibitor KPT-8602 Ameliorates Parkinson′s Disease by Inhibiting the NF-κB/NLRP3 Pathway. Front. Pharmacol. 2022, 13, 847605. [Google Scholar] [CrossRef] [PubMed]
  255. Li, Q.; Feng, H.; Wang, H.; Wang, Y.; Mou, W.; Xu, G.; Zhang, P.; Li, R.; Shi, W.; Wang, Z.; et al. Licochalcone B specifically inhibits the NLRP3 inflammasome by disrupting NEK7-NLRP3 interaction. EMBO Rep. 2022, 23, e53499. [Google Scholar] [CrossRef] [PubMed]
  256. Wang, H.R.; Tang, J.Y.; Wang, Y.Y.; Farooqi, A.A.; Yen, C.Y.; Yuan, S.F.; Huang, H.W.; Chang, H.W. Manoalide Preferentially Provides Antiproliferation of Oral Cancer Cells by Oxidative Stress-Mediated Apoptosis and DNA Damage. Cancers 2019, 11, 1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Folmer, F.; Jaspars, M.; Schumacher, M.; Dicato, M.; Diederich, M. Marine Natural Products Targeting Phospholipases A2. Biochem. Pharmacol. 2010, 80, 1793–1800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Salam, K.A.; Furuta, A.; Noda, N.; Tsuneda, S.; Sekiguchi, Y.; Yamashita, A.; Moriishi, K.; Nakakoshi, M.; Tsubuki, M.; Tani, H.; et al. Inhibition of Hepatitis C Virus NS3 Helicase by Manoalide. J. Nat. Prod. 2012, 75, 650–654. [Google Scholar] [CrossRef]
  259. Li, C.; Lin, H.; He, H.; Ma, M.; Jiang, W.; Zhou, R. Inhibition of the NLRP3 Inflammasome Activation by Manoalide Ameliorates Experimental Autoimmune Encephalomyelitis Pathogenesis. Front. Cell Dev. Biol. 2022, 10, 822236. [Google Scholar] [CrossRef]
  260. Fu, Q.; Li, J.; Qiu, L.; Ruan, J.; Mao, M.; Li, S.; Mao, Q. Inhibiting NLRP3 inflammasome with MCC950 ameliorates perioperative neurocognitive disorders, suppressing neuroinflammation in the hippocampus in aged mice. Int. Immunopharmacol. 2020, 82, 106317. [Google Scholar] [CrossRef]
  261. Swanton, T.; Beswick, J.A.; Hammadi, H.; Morris, L.; Williams, D.; de Cesco, S.; El-Sharkawy, L.; Yu, S.; Green, J.; Davis, J.B.; et al. Selective inhibition of the K+ efflux sensitive NLRP3 pathway by Cl- channel modulation. Chem. Sci. 2020, 11, 11720–11728. [Google Scholar] [CrossRef]
  262. Muñoz-Jurado, A.; Escribano, B.M.; Caballero-Villarraso, J.; Galván, A.; Agüera, E.; Santamaría, A.; Túnez, I. Melatonin and multiple sclerosis: Antioxidant, anti-inflammatory and immunomodulator mechanism of action. Inflammopharmacology 2022, 5, 1569–1596. [Google Scholar] [CrossRef]
  263. Madhu, L.N.; Kodali, M.; Attaluri, S.; Shuai, B.; Melissari, L.; Rao, X.; Shetty, A.K. Melatonin improves brain function in a model of chronic Gulf War Illness with modulation of oxidative stress, NLRP3 inflammasomes, and BDNF-ERK-CREB pathway in the hippocampus. Redox Biol. 2021, 43, 101973. [Google Scholar] [CrossRef]
  264. Fan, L.; Zhaohong, X.; Xiangxue, W.; Yingying, X.; Xiao, Z.; Xiaoyan, Z.; Jieke, Y.; Chao, L. Melatonin Ameliorates the Progression of Alzheimer′s Disease by Inducing TFEB Nuclear Translocation, Promoting Mitophagy, and Regulating NLRP3 Inflammasome Activity. Biomed. Res. Int. 2022, 2022, 8099459. [Google Scholar] [CrossRef]
  265. Farré-Alins, V.; Narros-Fernández, P.; Palomino-Antolín, A.; Decouty-Pérez, C.; Lopez-Rodriguez, A.B.; Parada, E.; Muñoz-Montero, A.; Gómez-Rangel, V.; López-Muñoz, F.; Ramos, E.; et al. Melatonin Reduces NLRP3 Inflammasome Activation by Increasing α7 nAChR-Mediated Autophagic Flux. Antioxidants 2020, 9, 1299. [Google Scholar] [CrossRef]
  266. Zheng, R.; Ruan, Y.; Yan, Y.; Lin, Z.; Xue, N.; Yan, Y.; Tian, J.; Yin, X.; Pu, J.; Zhang, B. Melatonin Attenuates Neuroinflammation by Down-Regulating NLRP3 Inflammasome via a SIRT1-Dependent Pathway in MPTP-Induced Models of Parkinson′s Disease. J. Inflamm. Res. 2021, 14, 3063–3075. [Google Scholar] [CrossRef]
  267. Zhang, Y.; Zhang, H.; Li, S.; Huang, K.; Jiang, L.; Wang, Y. Metformin Alleviates LPS-Induced Acute Lung Injury by Regulating the SIRT1/NF-κB/NLRP3 Pathway and Inhibiting Endothelial Cell Pyroptosis. Front. Pharmacol. 2022, 13, 801337. [Google Scholar] [CrossRef]
  268. Chen, Q.; Yin, Y.; Li, L.; Zhang, Y.; He, W.; Shi, Y. Milrinone Ameliorates the Neuroinflammation and Memory Function of Alzheimer′s Disease in an APP/PS1 Mouse Model. Neuropsychiatr. Dis. Treat. 2021, 17, 2129–2139. [Google Scholar] [CrossRef]
  269. Garcez, M.L.; Mina, F.; Bellettini-Santos, T.; da Luz, A.P.; Schiavo, G.L.; Macieski, J.M.C.; Medeiros, E.B.; Marques, A.O.; Magnus, N.Q.; Budni, J. The Involvement of NLRP3 on the Effects of Minocycline in an AD-Like Pathology Induced by β-Amyloid Oligomers Administered to Mice. Mol. Neurobiol. 2019, 56, 2606–2617. [Google Scholar] [CrossRef]
  270. Cruz, S.L.; Armenta-Reséndiz, M.; Carranza-Aguilar, C.J.; Galván, E.J. Minocycline prevents neuronal hyperexcitability and neuroinflammation in medial prefrontal cortex, as well as memory impairment caused by repeated toluene inhalation in adolescent rats. Toxicol. Appl. Pharmacol. 2020, 395, 114980. [Google Scholar] [CrossRef]
  271. Chen, W.; Guo, C.; Huang, S.; Jia, Z.; Wang, J.; Zhong, J.; Ge, H.; Yuan, J.; Chen, T.; Liu, X.; et al. MitoQ attenuates brain damage by polarizing microglia towards the M2 phenotype through inhibition of the NLRP3 inflammasome after ICH. Pharmacol. Res. 2020, 161, 105122. [Google Scholar] [CrossRef]
  272. Chen, W.; Teng, X.; Ding, H.; Xie, Z.; Cheng, P.; Liu, Z.; Feng, T.; Zhang, X.; Huang, W.; Geng, D. Nrf2 protects against cerebral ischemia-reperfusion injury by suppressing programmed necrosis and inflammatory signaling pathways. Ann. Transl. Med. 2022, 10, 285. [Google Scholar] [CrossRef]
  273. Li, C.; Wang, J.; Fang, Y.; Liu, Y.; Chen, T.; Sun, H.; Zhou, X.F.; Liao, H. Nafamostat mesilate improves function recovery after stroke by inhibiting neuroinflammation in rats. Brain Behav. Immun. 2016, 56, 230–245. [Google Scholar] [CrossRef]
  274. Coll, R.C.; Schroder, K.; Pelegrín, P. NLRP3 and pyroptosis blockers for treating inflammatory diseases. Trends Pharmacol. Sci. 2022, 43, 653–668. [Google Scholar] [CrossRef] [PubMed]
  275. Zhang, X.; Xu, A.; Ran, Y.; Wei, C.; Xie, F.; Wu, J. Design, synthesis, and biological evaluation of phenyl vinyl sulfone based NLRP3 inflammasome inhibitors. Bioorg. Chem. 2022, 128, 106010. [Google Scholar] [CrossRef] [PubMed]
  276. Wang, J.; Zheng, B.; Yang, S.; Tang, X.; Wang, J.; Wei, D. The protective effects of phoenixin-14 against lipopolysaccharide-induced inflammation and inflammasome activation in astrocytes. Inflamm. Res. 2020, 69, 779–787. [Google Scholar] [CrossRef] [PubMed]
  277. Dong, A.Q.; Yang, Y.P.; Jiang, S.M.; Yao, X.Y.; Qi, D.; Mao, C.J.; Cheng, X.Y.; Wang, F.; Hu, L.F.; Liu, C.F. Pramipexole inhibits astrocytic NLRP3 inflammasome activation via Drd3-dependent autophagy in a mouse model of Parkinson′s disease. Acta Pharmacol. Sin. 2022, 44, 32–43. [Google Scholar] [CrossRef]
  278. Yu, H.; Wu, M.; Lu, G.; Cao, T.; Chen, N.; Zhang, Y.; Jiang, Z.; Fan, H.; Yao, R. Prednisone alleviates demyelination through regulation of the NLRP3 inflammasome in a C57BL/6 mouse model of cuprizone-induced demyelination. Brain Res. 2018, 1678, 75–84. [Google Scholar] [CrossRef]
  279. Wei, C.; Guo, S.; Liu, W.; Jin, F.; Wei, B.; Fan, H.; Su, H.; Liu, J.; Zhang, N.; Fang, D.; et al. Resolvin D1 ameliorates Inflammation-Mediated Blood-Brain Barrier Disruption After Subarachnoid Hemorrhage in rats by Modulating A20 and NLRP3 Inflammasome. Front. Pharmacol. 2021, 11, 610734. [Google Scholar] [CrossRef]
  280. Zhang, J.; Guo, J.; Zhao, X.; Chen, Z.; Wang, G.; Liu, A.; Wang, Q.; Zhou, W.; Xu, Y.; Wang, C. Phosphodiesterase-5 inhibitor sildenafil prevents neuroinflammation, lowers beta-amyloid levels and improves cognitive performance in APP/PS1 transgenic mice. Behav. Brain Res. 2013, 250, 230–237. [Google Scholar] [CrossRef]
  281. Liu, Y.; Dai, Y.; Li, Q.; Chen, C.; Chen, H.; Song, Y.; Hua, F.; Zhang, Z. Beta-amyloid activates NLRP3 inflammasome via TLR4 in mouse microglia. Neurosci. Lett. 2020, 736, 135279. [Google Scholar] [CrossRef]
  282. Liao, Y.J.; Pan, R.Y.; Kong, X.X.; Cheng, Y.; Du, L.; Wang, Z.C.; Yuan, C.; Cheng, J.B.; Yuan, Z.Q.; Zhang, H.Y. Correction: 1,2,4-Trimethoxybenzene selectively inhibits NLRP3 inflammasome activation and attenuates experimental autoimmune encephalomyelitis. Acta Pharmacol. Sin. 2022, 43, 504, Erratum in Acta Pharmacol. Sin. 2020, 42, 1769–1779. [Google Scholar] [CrossRef]
  283. Qiu, J.; Chen, Y.; Zhuo, J.; Zhang, L.; Liu, J.; Wang, B.; Sun, D.; Yu, S.; Lou, H. Urolithin A promotes mitophagy and suppresses NLRP3 inflammasome activation in lipopolysaccharide-induced BV2 microglial cells and MPTP-induced Parkinson′s disease model. Neuropharmacology 2022, 207, 108963. [Google Scholar] [CrossRef]
  284. Tian, D.; Xing, Y.; Gao, W.; Zhang, H.; Song, Y.; Tian, Y.; Dai, Z. Sevoflurane Aggravates the Progress of Alzheimer′s Disease Through NLRP3/Caspase-1/Gasdermin D Pathway. Front. Cell Dev. Biol. 2022, 9, 801422. [Google Scholar] [CrossRef]
  285. Das, S.; Mishra, K.P.; Ganju, L.; Singh, S.B. Andrographolide - A promising therapeutic agent, negatively regulates glial cell derived neurodegeneration of prefrontal cortex, hippocampus and working memory impairment. J. Neuroimmunol. 2017, 313, 161–175. [Google Scholar] [CrossRef]
  286. Gugliandolo, E.; D′Amico, R.; Cordaro, M.; Fusco, R.; Siracusa, R.; Crupi, R.; Impellizzeri, D.; Cuzzocrea, S.; Di Paola, R. Neuroprotective Effect of Artesunate in Experimental Model of Traumatic Brain Injury. Front. Neurol. 2018, 9, 590. [Google Scholar] [CrossRef] [Green Version]
  287. Ju, I.G.; Huh, E.; Kim, N.; Lee, S.; Choi, J.G.; Hong, J.; Oh, M.S. Artemisiae Iwayomogii Herba inhibits lipopolysaccharide-induced neuroinflammation by regulating NF-κB and MAPK signaling pathways. Phytomedicine 2021, 84, 153501. [Google Scholar] [CrossRef]
  288. Li, M.; Li, H.; Fang, F.; Deng, X.; Ma, S. Astragaloside IV attenuates cognitive impairments induced by transient cerebral ischemia and reperfusion in mice via anti-inflammatory mechanisms. Neurosci. Lett. 2017, 639, 114–119. [Google Scholar] [CrossRef]
  289. Jin, X.; Liu, M.Y.; Zhang, D.F.; Zhong, X.; Du, K.; Qian, P.; Yao, W.F.; Gao, H.; Wei, M.J. Baicalin mitigates cognitive impairment and protects neurons from microglia-mediated neuroinflammation via suppressing NLRP3 inflammasomes and TLR4/NF-κB signaling pathway. CNS Neurosci. Ther. 2019, 25, 575–590. [Google Scholar] [CrossRef]
  290. Lee, C.M.; Lee, D.S.; Jung, W.K.; Yoo, J.S.; Yim, M.J.; Choi, Y.H.; Park, S.; Seo, S.K.; Choi, J.S.; Lee, Y.M.; et al. Benzyl isothiocyanate inhibits inflammasome activation in E. coli LPS-stimulated BV2 cells. Int. J. Mol. Med. 2016, 38, 912–918. [Google Scholar] [CrossRef] [Green Version]
  291. Yu, Y.; Wu, D.M.; Li, J.; Deng, S.H.; Liu, T.; Zhang, T.; He, M.; Zhao, Y.Y.; Xu, Y. Bixin Attenuates Experimental Autoimmune Encephalomyelitis by Suppressing TXNIP/NLRP3 Inflammasome Activity and Activating NRF2 Signaling. Front. Immunol. 2020, 11, 593368. [Google Scholar] [CrossRef]
  292. Satoh, T.; Trudler, D.; Oh, C.K.; Lipton, S.A. Potential Therapeutic Use of the Rosemary Diterpene Carnosic Acid for Alzheimer′s Disease, Parkinson′s Disease, and Long-COVID through NRF2 Activation to Counteract the NLRP3 Inflammasome. Antioxidants 2022, 11, 124. [Google Scholar] [CrossRef]
  293. Shi, W.; Xu, G.; Zhan, X.; Gao, Y.; Wang, Z.; Fu, S.; Qin, N.; Hou, X.; Ai, Y.; Wang, C.; et al. Carnosol inhibits inflammasome activation by directly targeting HSP90 to treat inflammasome-mediated diseases. Cell Death Dis. 2020, 11, 252. [Google Scholar] [CrossRef] [Green Version]
  294. Chu, X.; Zhang, L.; Zhou, Y.; Fang, Q. Cucurbitacin B alleviates cerebral ischemia/reperfusion injury by inhibiting NLRP3 inflammasome-mediated inflammation and reducing oxidative stress. Biosci. Biotechnol. Biochem. 2022, 11, zbac065. [Google Scholar] [CrossRef] [PubMed]
  295. González-Cofrade, L.; Cuadrado, I.; Amesty, Á.; Estévez-Braun, A.; de Las Heras, B.; Hortelano, S. Dehydroisohispanolone as a Promising NLRP3 Inhibitor Agent: Bioevaluation and Molecular Docking. Pharmaceuticals 2022, 15, 825. [Google Scholar] [CrossRef] [PubMed]
  296. Zhang, Y.; Liu, D.; Yao, X.; Wen, J.; Wang, Y.; Zhang, Y. DMTHB ameliorates memory impairment in Alzheimer′s disease mice through regulation of neuroinflammation. Neurosci. Lett. 2022, 785, 136770. [Google Scholar] [CrossRef] [PubMed]
  297. Yang, H.; Chen, Y.; Yu, L.; Xu, Y. Esculentoside A exerts anti-inflammatory activity in microglial cells. Int. Immunopharmacol. 2017, 51, 148–157. [Google Scholar] [CrossRef]
  298. Zheng, X.; Gong, T.; Tang, C.; Zhong, Y.; Shi, L.; Fang, X.; Chen, D.; Zhu, Z. Gastrodin improves neuroinflammation-induced cognitive dysfunction in rats by regulating NLRP3 inflammasome. BMC Anesthesiol. 2022, 22, 371. [Google Scholar] [CrossRef]
  299. Zhang, Y.; Zhao, Y.; Zhang, J.; Gao, Y.; Li, S.; Chang, C.; Yu, D.; Yang, G. Ginkgolide B inhibits NLRP3 inflammasome activation and promotes microglial M2 polarization in Aβ1-42-induced microglia cells. Neurosci. Lett. 2021, 764, 136206. [Google Scholar] [CrossRef]
  300. Shao, L.; Dong, C.; Geng, D.; He, Q.; Shi, Y. Ginkgolide B inactivates the NLRP3 inflammasome by promoting autophagic degradation to improve learning and memory impairment in Alzheimer′s disease. Metab. Brain Dis. 2022, 37, 329–341. [Google Scholar] [CrossRef]
  301. Jiang, J.; Sun, X.; Akther, M.; Lian, M.L.; Quan, L.H.; Koppula, S.; Han, J.H.; Kopalli, S.R.; Kang, T.B.; Lee, K.H. Ginsenoside metabolite 20(S)-protopanaxatriol from Panax ginseng attenuates inflammation-mediated NLRP3 inflammasome activation. J. Ethnopharmacol. 2020, 251, 112564. [Google Scholar] [CrossRef]
  302. Gao, Y.; Li, J.; Wang, J.; Li, X.; Li, J.; Chu, S.; Li, L.; Chen, N.; Zhang, L. Ginsenoside Rg1 prevent and treat inflammatory diseases: A review. Int. Immunopharmacol. 2020, 87, 106805. [Google Scholar] [CrossRef]
  303. Wang, J.; Wang, D.; Zhou, Z.; Zhang, X.; Zhang, C.; He, Y.; Liu, C.; Yuan, C.; Yuan, D.; Wang, T. Saponins from Panax japonicus alleviate HFD-induced impaired behaviors through inhibiting NLRP3 inflammasome to upregulate AMPA receptors. Neurochem. Int. 2021, 148, 105098. [Google Scholar] [CrossRef]
  304. Yi, Y.S. Roles of ginsenosides in inflammasome activation. J. Ginseng Res. 2019, 43, 172–178. [Google Scholar] [CrossRef]
  305. Yi, Y.S. New mechanisms of ginseng saponin-mediated anti-inflammatory action via targeting canonical inflammasome signaling pathways. J. Ethnopharmacol. 2021, 278, 114292. [Google Scholar] [CrossRef]
  306. Chaturvedi, S.; Tiwari, V.; Gangadhar, N.M.; Rashid, M.; Sultana, N.; Singh, S.K.; Shukla, S.; Wahajuddin, M. Isoformononetin, a dietary isoflavone protects against streptozotocin induced rat model of neuroinflammation through inhibition of NLRP3/ASC/IL-1 axis activation. Life Sci. 2021, 286, 119989. [Google Scholar] [CrossRef]
  307. Zeng, J.; Chen, Y.; Ding, R. Isoliquiritigenin alleviates early brain injury after experimental intracerebral hemorrhage via suppressing ROS- and/or NF-κB-mediated NLRP3 inflammasome activation by promoting Nrf2 antioxidant pathway. J. Neuroinflamm. 2017, 14, 119. [Google Scholar] [CrossRef] [Green Version]
  308. Wang, Y.H.; Lv, H.N.; Cui, Q.H.; Tu, P.F.; Jiang, Y.; Zeng, K.W. Isosibiricin inhibits microglial activation by targeting the dopamine D1/D2 receptor-dependent NLRP3/caspase-1 inflammasome pathway. Acta Pharmacol. Sin. 2020, 41, 173–180. [Google Scholar] [CrossRef]
  309. Manach, C.; Scalbert, A.; Morand, C.; Rémésy, C.; Jiménez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [CrossRef] [Green Version]
  310. Nabavi, S.F.; Sureda, A.; Dehpour, A.R.; Shirooie, S.; Silva, A.S.; Devi, K.P.; Ahmed, T.; Ishaq, N.; Hashim, R.; Sobarzo-Sánchez, E.; et al. Regulation of autophagy by polyphenols: Paving the road for treatment of neurodegeneration. Biotechnol. Adv. 2018, 36, 1768–1778. [Google Scholar] [CrossRef]
  311. Han, X.; Sun, S.; Sun, Y.; Song, Q.; Zhu, J.; Song, N.; Chen, M.; Sun, T.; Xia, M.; Ding, J.; et al. Small molecule-driven NLRP3 inflammation inhibition via interplay between ubiquitination and autophagy: Implications for Parkinson disease. Autophagy 2019, 15, 1860–1881. [Google Scholar] [CrossRef]
  312. Mokarizadeh, N.; Karimi, P.; Erfani, M.; Sadigh-Eteghad, S.; Fathi Maroufi, N.; Rashtchizadeh, N. β-Lapachone attenuates cognitive impairment and neuroinflammation in beta-amyloid induced mouse model of Alzheimer′s disease. Int. Immunopharmacol. 2020, 81, 106300. [Google Scholar] [CrossRef]
  313. Xiong, R.; Zhou, X.G.; Tang, Y.; Wu, J.M.; Sun, Y.S.; Teng, J.F.; Pan, R.; Law, B.Y.; Zhao, Y.; Qiu, W.Q.; et al. Lychee seed polyphenol protects the blood-brain barrier through inhibiting Aβ(25-35)-induced NLRP3 inflammasome activation via the AMPK/mTOR/ULK1-mediated autophagy in bEnd.3 cells and APP/PS1 mice. Phytother. Res. 2021, 35, 954–973. [Google Scholar] [CrossRef]
  314. Qiu, W.Q.; Pan, R.; Tang, Y.; Zhou, X.G.; Wu, J.M.; Yu, L.; Law, B.Y.; Ai, W.; Yu, C.L.; Qin, D.L.; et al. Lychee seed polyphenol inhibits Aβ-induced activation of NLRP3 inflammasome via the LRP1/AMPK mediated autophagy induction. Biomed. Pharmacother. 2020, 130, 110575. [Google Scholar] [CrossRef] [PubMed]
  315. Lei, L.Y.; Wang, R.C.; Pan, Y.L.; Yue, Z.G.; Zhou, R.; Xie, P.; Tang, Z.S. Mangiferin inhibited neuroinflammation through regulating microglial polarization and suppressing NF-κB, NLRP3 pathway. Chin. J. Nat. Med. 2021, 19, 112–119. [Google Scholar] [CrossRef] [PubMed]
  316. Gong, J.; Luo, S.; Zhao, S.; Yin, S.; Li, X.; Mou, T. Myricitrin attenuates memory impairment in a rat model of sepsis-associated encephalopathy via the NLRP3/Bax/Bcl pathway. Folia Neuropathol. 2019, 57, 327–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Yan, T.; Lu, M. Myricitrin attenuates hypoxic-ischemia-induced brain injury in neonatal rats by mitigating oxidative stress and nuclear factor erythroid 2-related factor 2/hemeoxygenase-1/antioxidant response element signaling pathway. Phcog. Mag. 2021, 17, 828–835. [Google Scholar] [CrossRef]
  318. Wang, H.; Guo, Y.; Qiao, Y.; Zhang, J.; Jiang, P. Nobiletin Ameliorates NLRP3 Inflammasome-Mediated Inflammation Through Promoting Autophagy via the AMPK Pathway. Mol. Neurobiol. 2020, 57, 5056–5068. [Google Scholar] [CrossRef]
  319. Al Rihani, S.B.; Darakjian, L.I.; Kaddoumi, A. Oleocanthal-Rich Extra-Virgin Olive Oil Restores the Blood-Brain Barrier Function through NLRP3 Inflammasome Inhibition Simultaneously with Autophagy Induction in TgSwDI Mice. ACS Chem. Neurosci. 2019, 10, 3543–3554. [Google Scholar] [CrossRef]
  320. Wang, S.; Yang, H.; Yu, L.; Jin, J.; Qian, L.; Zhao, H.; Xu, Y.; Zhu, X. Oridonin attenuates Aβ1-42-induced neuroinflammation and inhibits NF-κB pathway. PloS ONE 2014, 9, e104745. [Google Scholar] [CrossRef] [Green Version]
  321. Liu, Y.; Chen, X.; Gong, Q.; Shi, J.; Li, F. Osthole Improves Cognitive Function of Vascular Dementia Rats: Reducing Aβ Deposition via Inhibition NLRP3 Inflammasome. Biol. Pharm. Bull. 2020, 43, 1315–1323. [Google Scholar] [CrossRef]
  322. Chen, D.B.; Gao, H.W.; Peng, C.; Pei, S.Q.; Dai, A.R.; Yu, X.T.; Zhou, P.; Wang, Y.; Cai, B. Quinones as preventive agents in Alzheimer′s diseases: Focus on NLRP3 inflammasomes. J. Pharm. Pharmacol. 2020, 72, 1481–1490. [Google Scholar] [CrossRef]
  323. Han, X.; Xu, T.; Fang, Q.; Zhang, H.; Yue, L.; Hu, G.; Sun, L. Quercetin hinders microglial activation to alleviate neurotoxicity via the interplay between NLRP3 inflammasome and mitophagy. Redox Biol. 2021, 44, 102010. [Google Scholar] [CrossRef]
  324. Li, H.; Chen, F.J.; Yang, W.L.; Qiao, H.Z.; Zhang, S.J. Quercetin improves cognitive disorder in aging mice by inhibiting NLRP3 inflammasome activation. Food Funct. 2021, 12, 717–725. [Google Scholar] [CrossRef]
  325. Kiasalari, Z.; Afshin-Majd, S.; Baluchnejadmojarad, T.; Azadi-Ahmadabadi, E.; Fakour, M.; Ghasemi-Tarie, R.; Jalalzade-Ogvar, S.; Khodashenas, V.; Tashakori-Miyanroudi, M.; Roghani, M. Sinomenine Alleviates Murine Experimental Autoimmune Encephalomyelitis Model of Multiple Sclerosis through Inhibiting NLRP3 Inflammasome. J. Mol. Neurosci. 2021, 71, 215–224. [Google Scholar] [CrossRef]
  326. Atluri, V.S.R.; Tiwari, S.; Rodriguez, M.; Kaushik, A.; Yndart, A.; Kolishetti, N.; Yatham, M.; Nair, M. Inhibition of Amyloid-Beta Production, Associated Neuroinflammation, and Histone Deacetylase 2-Mediated Epigenetic Modifications Prevent Neuropathology in Alzheimer′s Disease in vitro Model. Front. Aging Neurosci. 2020, 11, 342. [Google Scholar] [CrossRef] [Green Version]
  327. Cadoná, F.C.; de Souza, D.V.; Fontana, T.; Bodenstein, D.F.; Ramos, A.P.; Sagrillo, M.R.; Salvador, M.; Mota, K.; Davidson, C.B.; Ribeiro, E.E.; et al. Açaí (Euterpe oleracea Mart.) as a Potential Anti-neuroinflammatory Agent: NLRP3 Priming and Activating Signal Pathway Modulation. Mol. Neurobiol. 2021, 58, 4460–4476. [Google Scholar] [CrossRef]
  328. Yu, S.H.; Sun, X.; Kim, M.K.; Akther, M.; Han, J.H.; Kim, T.Y.; Jiang, J.; Kang, T.B.; Lee, K.H. Chrysanthemum indicum extract inhibits NLRP3 and AIM2 inflammasome activation via regulating ASC phosphorylation. J. Ethnopharmacol. 2019, 239, 111917. [Google Scholar] [CrossRef]
  329. Jeong, Y.H.; Kim, T.I.; Oh, Y.C.; Ma, J.Y. Chrysanthemum indicum Prevents Hydrogen Peroxide-Induced Neurotoxicity by Activating the TrkB/Akt Signaling Pathway in Hippocampal Neuronal Cells. Nutrients 2021, 13, 3690. [Google Scholar] [CrossRef]
  330. Wang, Z.; Xu, G.; Li, Z.; Xiao, X.; Tang, J.; Bai, Z. NLRP3 Inflammasome Pharmacological Inhibitors in Glycyrrhiza for NLRP3-Driven Diseases Treatment: Extinguishing the Fire of Inflammation. J. Inflamm. Res. 2022, 15, 409–422. [Google Scholar] [CrossRef]
  331. Kim, N.; Do, J.; Ju, I.G.; Jeon, S.H.; Lee, J.K.; Oh, M.S. Picrorhiza kurroa Prevents Memory Deficits by Inhibiting NLRP3 Inflammasome Activation and BACE1 Expression in 5xFAD Mice. Neurotherapeutics 2020, 17, 189–199. [Google Scholar] [CrossRef]
  332. Huang, Z.; Zhou, X.; Zhang, X.; Huang, L.; Sun, Y.; Cheng, Z.; Xu, W.; Li, C.G.; Zheng, Y.; Huang, M. Pien-Tze-Huang, a Chinese patent formula, attenuates NLRP3 inflammasome-related neuroinflammation by enhancing autophagy via the AMPK/mTOR/ULK1 signaling pathway. Biomed. Pharmacother. 2021, 141, 111814. [Google Scholar] [CrossRef]
  333. Yin, X.L.; Wu, H.M.; Zhang, B.H.; Zhu, N.W.; Chen, T.; Ma, X.X.; Zhang, L.Y.; Lv, L.; Zhang, M.; Wang, F.Y.; et al. Tojapride prevents CaSR-mediated NLRP3 inflammasome activation in oesophageal epithelium irritated by acidic bile salts. J. Cell. Mol. Med. 2020, 24, 1208–1219. [Google Scholar] [CrossRef] [Green Version]
  334. Zhu, T.; Fang, B.Y.; Meng, X.B.; Zhang, S.X.; Wang, H.; Gao, G.; Liu, F.; Wu, Y.; Hu, J.; Sun, G.B.; et al. Folium Ginkgo extract and tetramethylpyrazine sodium chloride injection (Xingxiong injection) protects against focal cerebral ischaemia/reperfusion injury via activating the Akt/Nrf2 pathway and inhibiting NLRP3 inflammasome activation. Pharm. Biol. 2022, 60, 195–205. [Google Scholar] [CrossRef] [PubMed]
  335. Kim, H.; Hong, J.Y.; Jeon, W.J.; Lee, J.; Baek, S.H.; Ha, I.H. Lycopus lucidus Turcz Exerts Neuroprotective Effects Against H2O2-Induced Neuroinflammation by Inhibiting NLRP3 Inflammasome Activation in Cortical Neurons. J. Inflamm. Res. 2021, 14, 1759–1773. [Google Scholar] [CrossRef]
  336. Denes, A.; Coutts, G.; Lénárt, N.; Cruickshank, S.M.; Pelegrin, P.; Skinner, J.; Rothwell, N.; Allan, S.M.; Brough, D. AIM2 and NLRC4 inflammasomes contribute with ASC to acute brain injury independently of NLRP3. Proc. Natl. Acad. Sci. USA 2015, 112, 4050–4055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Yang, X.L.; Wang, X.; Shao, L.; Jiang, G.T.; Min, J.W.; Mei, X.Y.; He, X.H.; Liu, W.H.; Huang, W.X.; Peng, B.W. TRPV1 mediates astrocyte activation and interleukin-1β release induced by hypoxic ischemia (HI). J. Neuroinflamm. 2019, 16, 114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  338. Schölwer, I.; Habib, P.; Voelz, C.; Rolfes, L.; Beyer, C.; Slowik, A. NLRP3 Depletion Fails to Mitigate Inflammation but Restores Diminished Phagocytosis in BV-2 Cells After In Vitro Hypoxia. Mol. Neurobiol. 2020, 57, 2588–2599. [Google Scholar] [CrossRef]
  339. Sun, Y.; Ma, J.; Li, D.; Li, P.; Zhou, X.; Li, Y.; He, Z.; Qin, L.; Liang, L.; Luo, X. Interleukin-10 inhibits interleukin-1β production and inflammasome activation of microglia in epileptic seizures. J. Neuroinflamm. 2019, 16, 66. [Google Scholar] [CrossRef] [Green Version]
  340. Liew, F.; Girard, J.P.; Turnquist, H. Interleukin-33 in health and disease. Nat. Rev. Immunol. 2016, 16, 676–689. [Google Scholar] [CrossRef]
  341. Jiao, M.; Li, X.; Chen, L.; Wang, X.; Yuan, B.; Liu, T.; Dong, Q.; Mei, H.; Yin, H. Neuroprotective effect of astrocyte-derived IL-33 in neonatal hypoxic-ischemic brain injury. J. Neuroinflamm. 2020, 17, 251. [Google Scholar] [CrossRef]
  342. Strangward, P.; Haley, M.J.; Albornoz, M.G.; Barrington, J.; Shaw, T.; Dookie, R.; Zeef, L.; Baker, S.M.; Winter, E.; Tzeng, T.C.; et al. Targeting the IL33-NLRP3 axis improves therapy for experimental cerebral malaria. Proc. Natl. Acad. Sci. USA 2018, 115, 7404–7409. [Google Scholar] [CrossRef] [Green Version]
  343. Bellut, M.; Raimondi, A.T.; Haarmann, A.; Zimmermann, L.; Stoll, G.; Schuhmann, M.K. NLRP3 Inhibition Reduces rt-PA Induced Endothelial Dysfunction under Ischemic Conditions. Biomedicines 2022, 10, 762. [Google Scholar] [CrossRef]
  344. Xu, Q.; Ye, Y.; Wang, Z.; Zhu, H.; Li, Y.; Wang, J.; Gao, W.; Gu, L. NLRP3 Knockout Protects against Lung Injury Induced by Cerebral Ischemia-Reperfusion. Oxid. Med. Cell. Longev. 2022, 2022, 6260102. [Google Scholar] [CrossRef]
  345. Chen, B.; Zhang, M.; Ji, M.; Zhang, D.; Chen, B.; Gong, W.; Li, X.; Zhou, Y.; Dong, C.; Wen, G.; et al. The neuroprotective mechanism of lithium after ischaemic stroke. Commun. Biol. 2022, 5, 105. [Google Scholar] [CrossRef]
  346. Zhang, Y.; Wang, Y.; Zhao, W.; Li, L.; Li, L.; Sun, Y.; Shao, J.; Ren, X.; Zang, W.; Cao, J. Role of spinal RIP3 in inflammatory pain and electroacupuncture-mediated analgesic effect in mice. Life Sci. 2022, 306, 120839. [Google Scholar] [CrossRef]
  347. Zhong, X.; Chen, B.; Li, Z.; Lin, R.; Ruan, S.; Wang, F.; Liang, H.; Tao, J. Electroacupuncture Ameliorates Cognitive Impairment Through the Inhibition of NLRP3 Inflammasome Activation by Regulating Melatonin-Mediated Mitophagy in Stroke Rats. Neurochem. Res. 2022, 47, 1917–1930, Erratum in Neurochem. Res. 2022, 47, 1931–1933. [Google Scholar] [CrossRef]
  348. Halle, A.; Hornung, V.; Petzold, G.C.; Stewart, C.R.; Monks, B.G.; Reinheckel, T.; Fitzgerald, K.A.; Latz, E.; Moore, K.J.; Golenbock, D.T. The NALP3 inflammasome is involved in the innate immune response to amyloid-beta. Nat. Immunol. 2008, 9, 857–865. [Google Scholar] [CrossRef] [Green Version]
  349. Heneka, M.T.; Kummer, M.P.; Stutz, A.; Delekate, A.; Schwartz, S.; Vieira-Saecker, A.; Griep, A.; Axt, D.; Remus, A.; Tzeng, T.C.; et al. NLRP3 is activated in Alzheimer′s disease and contributes to pathology in APP/PS1 mice. Nature 2013, 493, 674–678. [Google Scholar] [CrossRef] [Green Version]
  350. Lučiūnaitė, A.; McManus, R.M.; Jankunec, M.; Rácz, I.; Dansokho, C.; Dalgėdienė, I.; Schwartz, S.; Brosseron, F.; Heneka, M.T. Soluble Aβ oligomers and protofibrils induce NLRP3 inflammasome activation in microglia. J. Neurochem. 2020, 155, 650–661. [Google Scholar] [CrossRef] [Green Version]
  351. Couturier, J.; Stancu, I.C.; Schakman, O.; Pierrot, N.; Huaux, F.; Kienlen-Campard, P.; Dewachter, I.; Octave, J.N. Activation of phagocytic activity in astrocytes by reduced expression of the inflammasome component ASC and its implication in a mouse model of Alzheimer disease. J. Neuroinflamm. 2016, 13, 20. [Google Scholar] [CrossRef] [Green Version]
  352. Ramaswamy, S.B.; Bhagavan, S.M.; Kaur, H.; Giler, G.E.; Kempuraj, D.; Thangavel, R.; Ahmed, M.E.; Selvakumar, G.P.; Raikwar, S.P.; Zaheer, S.; et al. Glia Maturation Factor in the Pathogenesis of Alzheimer′s disease. Open Access J. Neurol. Neurosurg. 2019, 12, 79–82. [Google Scholar]
  353. Friker, L.L.; Scheiblich, H.; Hochheiser, I.V.; Brinkschulte, R.; Riedel, D.; Latz, E.; Geyer, M.; Heneka, M.T. β-Amyloid Clustering around ASC Fibrils Boosts Its Toxicity in Microglia. Cell Rep. 2020, 30, 3743–3754.e6. [Google Scholar] [CrossRef]
  354. Murphy, N.; Grehan, B.; Lynch, M.A. Glial uptake of amyloid beta induces NLRP3 inflammasome formation via cathepsin-dependent degradation of NLRP10. Neuromolecular Med. 2014, 16, 205–215. [Google Scholar] [CrossRef] [PubMed]
  355. Slowik, A.; Lammerding, L.; Zendedel, A.; Habib, P.; Beyer, C. Impact of steroid hormones E2 and P on the NLRP3/ASC/Casp1 axis in primary mouse astroglia and BV-2 cells after in vitro hypoxia. J. Steroid Biochem. Mol. Biol. 2018, 183, 18–26. [Google Scholar] [CrossRef] [PubMed]
  356. Hong, Y.; Liu, Y.; Yu, D.; Wang, M.; Hou, Y. The neuroprotection of progesterone against Aβ-induced NLRP3-Caspase-1 inflammasome activation via enhancing autophagy in astrocytes. Int. Immunopharmacol. 2019, 74, 105669. [Google Scholar] [CrossRef] [PubMed]
  357. Chen, D.; Dixon, B.J.; Doycheva, D.M.; Li, B.; Zhang, Y.; Hu, Q.; He, Y.; Guo, Z.; Nowrangi, D.; Flores, J.; et al. IRE1α inhibition decreased TXNIP/NLRP3 inflammasome activation through miR-17-5p after neonatal hypoxic-ischemic brain injury in rats. J. Neuroinflamm. 2018, 15, 32. [Google Scholar] [CrossRef] [Green Version]
  358. Wang, C.Y.; Xu, Y.; Wang, X.; Guo, C.; Wang, T.; Wang, Z.Y. Dl-3-n-Butylphthalide Inhibits NLRP3 Inflammasome and Mitigates Alzheimer′s-Like Pathology via Nrf2-TXNIP-TrX Axis. Antioxid. Redox. Signal. 2019, 30, 1411–1431. [Google Scholar] [CrossRef]
  359. Sun, Y.; Huang, J.; Chen, Y.; Shang, H.; Zhang, W.; Yu, J.; He, L.; Xing, C.; Zhuang, C. Direct inhibition of Keap1-Nrf2 Protein-Protein interaction as a potential therapeutic strategy for Alzheimer′s disease. Bioorg. Chem. 2020, 103, 104172. [Google Scholar] [CrossRef]
  360. Saad El-Din, S.; Rashed, L.; Medhat, E.; Emad Aboulhoda, B.; Desoky Badawy, A.; Mohammed ShamsEldeen, A.; Abdelgwad, M. Active form of vitamin D analogue mitigates neurodegenerative changes in Alzheimer′s disease in rats by targeting Keap1/Nrf2 and MAPK-38p/ERK signaling pathways. Steroids 2020, 156, 108586. [Google Scholar] [CrossRef]
  361. Yang, X.; Ji, J.; Liu, C.; Zhou, M.; Li, H.; Ye, S.; Hu, Q. HJ22, a Novel derivative of piperine, attenuates ibotenic acid-induced cognitive impairment, oxidative stress, apoptosis and inflammation via inhibiting the protein-protein interaction of Keap1-Nrf2. Int. Immunopharmacol. 2020, 83, 106383. [Google Scholar] [CrossRef]
  362. Yang, X.; Zhi, J.; Leng, H.; Chen, Y.; Gao, H.; Ma, J.; Ji, J.; Hu, Q. The piperine derivative HJ105 inhibits Aβ1-42-induced neuroinflammation and oxidative damage via the Keap1-Nrf2-TXNIP axis. Phytomedicine 2021, 87, 153571. [Google Scholar] [CrossRef]
  363. Bharti, V.; Tan, H.; Zhou, H.; Wang, J.F. Txnip mediates glucocorticoid-activated NLRP3 inflammatory signaling in mouse microglia. Neurochem. Int. 2019, 131, 104564. [Google Scholar] [CrossRef]
  364. Gussago, C.; Casati, M.; Ferri, E.; Arosio, B. The Triggering Receptor Expressed on Myeloid Cells-2 (TREM-2) as Expression of the Relationship between Microglia and Alzheimer′s Disease: A Novel Marker for a Promising Pathway to Explore. J. Frailty Aging 2019, 8, 54–56. [Google Scholar] [CrossRef]
  365. Wang, S.Y.; Gong, P.Y.; Yan, E.; Zhang, Y.D.; Jiang, T. The role of TREML2 in Alzheimer′s disease. J. Alzheimers Dis. 2020, 76, 799–806. [Google Scholar] [CrossRef]
  366. Sierksma, A.; Lu, A.; Mancuso, R.; Fattorelli, N.; Thrupp, N.; Salta, E.; Zoco, J.; Blum, D.; Buee, L.; De Strooper, B.; et al. Novel Alzheimer risk genes determine the microglia response to amyloid-beta but not to TAU pathology. EMBO Mol. Med. 2020, 12, e10606. [Google Scholar] [CrossRef]
  367. Zheng, H.; Liu, C.C.; Atagi, Y.; Chen, X.F.; Jia, L.; Yang, L.; He, W.; Zhang, X.; Kang, S.S.; Rosenberry, T.L.; et al. Opposing roles of the triggering receptor expressed on myeloid cells 2 and triggering receptor expressed on myeloid cells-like transcript 2 in microglia activation. Neurobiol. Aging 2016, 42, 132–141. [Google Scholar] [CrossRef] [Green Version]
  368. Wang, S.Y.; Fu, X.X.; Duan, R.; Wei, B.; Cao, H.M.; Yan, E.; Chen, S.Y.; Zhang, Y.D.; Jiang, T. The Alzheimer′s disease-associated gene TREML2 modulates inflammation by regulating microglia polarization and NLRP3 inflammasome activation. Neural Regen. Res. 2023, 18, 434–438. [Google Scholar] [CrossRef]
  369. Tejera, D.; Mercan, D.; Sanchez-Caro, J.M.; Hanan, M.; Greenberg, D.; Soreq, H.; Latz, E.; Golenbock, D.; Heneka, M.T. Systemic inflammation impairs microglial Aβ clearance through NLRP3 inflammasome. EMBO J. 2019, 38, e101064. [Google Scholar] [CrossRef]
  370. Lopez-Rodriguez, A.B.; Hennessy, E.; Murray, C.L.; Nazmi, A.; Delaney, H.J.; Healy, D.; Fagan, S.G.; Rooney, M.; Stewart, E.; Lewis, A.; et al. Acute systemic inflammation exacerbates neuroinflammation in Alzheimer′s disease: IL-1β drives amplified responses in primed astrocytes and neuronal network dysfunction. Alzheimers Dement. 2021, 17, 1735–1755. [Google Scholar] [CrossRef]
  371. Saresella, M.; Piancone, F.; Marventano, I.; Zoppis, M.; Hernis, A.; Zanette, M.; Trabattoni, D.; Chiappedi, M.; Ghezzo, A.; Canevini, M.P.; et al. Multiple inflammasome complexes are activated in autistic spectrum disorders. Brain Behav. Immun. 2016, 57, 125–133. [Google Scholar] [CrossRef]
  372. Ahmed, M.E.; Iyer, S.; Thangavel, R.; Kempuraj, D.; Selvakumar, G.P.; Raikwar, S.P.; Zaheer, S.; Zaheer, A. Co-Localization of Glia Maturation Factor with NLRP3 Inflammasome and Autophagosome Markers in Human Alzheimer′s Disease Brain. J. Alzheimers Dis. 2017, 60, 1143–1160. [Google Scholar] [CrossRef] [Green Version]
  373. Stancu, I.C.; Cremers, N.; Vanrusselt, H.; Couturier, J.; Vanoosthuyse, A.; Kessels, S.; Lodder, C.; Brône, B.; Huaux, F.; Octave, J.N.; et al. Aggregated Tau activates NLRP3-ASC inflammasome exacerbating exogenously seeded and non-exogenously seeded pathology in vivo. Acta Neuropathol. 2019, 137, 599–617. [Google Scholar] [CrossRef] [Green Version]
  374. Zhao, S.; Li, X.; Wang, J.; Wang, H. The Role of the Effects of Autophagy on NLRP3 Inflammasome in Inflammatory Nervous System Diseases. Front. Cell Dev. Biol. 2021, 9, 657478. [Google Scholar] [CrossRef] [PubMed]
  375. Zhou, W.; Xiao, D.; Zhao, Y.; Tan, B.; Long, Z.; Yu, L.; He, G. Enhanced Autolysosomal Function Ameliorates the Inflammatory Response Mediated by the NLRP3 Inflammasome in Alzheimer′s Disease. Front. Aging Neurosci. 2021, 13, 629891. [Google Scholar] [CrossRef] [PubMed]
  376. Zhang, P.; Shao, X.Y.; Qi, G.J.; Chen, Q.; Bu, L.L.; Chen, L.J.; Shi, J.; Ming, J.; Tian, B. Cdk5-Dependent Activation of Neuronal Inflammasomes in Parkinson′s Disease. Mov. Disord. 2016, 31, 366–376. [Google Scholar] [CrossRef] [PubMed]
  377. Deora, V.; Albornoz, E.A.; Zhu, K.; Woodruff, T.M.; Gordon, R. The Ketone Body β-Hydroxybutyrate Does Not Inhibit Synuclein Mediated Inflammasome Activation in Microglia. J. Neuroimmune Pharmacol. 2017, 12, 568–574. [Google Scholar] [CrossRef]
  378. Shippy, D.C.; Wilhelm, C.; Viharkumar, P.A.; Raife, T.J.; Ulland, T.K. β-Hydroxybutyrate inhibits inflammasome activation to attenuate Alzheimer′s disease pathology. J. Neuroinflamm. 2020, 17, 280. [Google Scholar] [CrossRef]
  379. Sarkar, S.; Malovic, E.; Harishchandra, D.S.; Ghaisas, S.; Panicker, N.; Charli, A.; Palanisamy, B.N.; Rokad, D.; Jin, H.; Anantharam, V.; et al. Mitochondrial impairment in microglia amplifies NLRP3 inflammasome proinflammatory signaling in cell culture and animal models of Parkinson′s disease. NPJ Park. Dis. 2017, 3, 30. [Google Scholar] [CrossRef] [Green Version]
  380. Gordon, R.; Albornoz, E.A.; Christie, D.C.; Langley, M.R.; Kumar, V.; Mantovani, S.; Robertson, A.A.B.; Butler, M.S.; Rowe, D.B.; O′Neill, L.A.; et al. Inflammasome Inhibition Prevents a-Synuclein Pathology and Dopaminergic Neurodegeneration in Mice. Sci. Transl. Med. 2018, 10, eaah4066. [Google Scholar] [CrossRef] [Green Version]
  381. Li, Y.; Xia, Y.; Yin, S.; Wan, F.; Hu, J.; Kou, L.; Sun, Y.; Wu, J.; Zhou, Q.; Huang, J.; et al. Targeting Microglial α-Synuclein/TLRs/NF-kappaB/NLRP3 Inflammasome Axis in Parkinson′s Disease. Front. Immunol. 2021, 12, 719807. [Google Scholar] [CrossRef]
  382. Scheiblich, H.; Bousset, L.; Schwartz, S.; Griep, A.; Latz, E.; Melki, R.; Heneka, M.T. Microglial NLRP3 Inflammasome Activation upon TLR2 and TLR5 Ligation by Distinct α-Synuclein Assemblies. J. Immunol. 2021, 207, 2143–2154. [Google Scholar] [CrossRef]
  383. von Herrmann, K.M.; Salas, L.A.; Martinez, E.M.; Young, A.L.; Howard, J.M.; Feldman, M.S.; Christensen, B.C.; Wilkins, O.M.; Lee, S.L.; Hickey, W.F.; et al. NLRP3 expression in mesencephalic neurons and characterization of a rare NLRP3 polymorphism associated with decreased risk of Parkinson′s disease. NPJ Park. Dis. 2018, 4, 24. [Google Scholar] [CrossRef]
  384. Fan, Z.; Pan, Y.T.; Zhang, Z.Y.; Yang, H.; Yu, S.Y.; Zheng, Y.; Ma, J.H.; Wang, X.M. Systemic Activation of NLRP3 Inflammasome and Plasma a-Synuclein Levels Are Correlated with Motor Severity and Progression in Parkinson′s Disease. J. Neuroinflamm. 2020, 17, 11. [Google Scholar] [CrossRef] [Green Version]
  385. Anderson, F.L.; von Herrmann, K.M.; Andrew, A.S.; Kuras, Y.I.; Young, A.L.; Scherzer, C.R.; Hickey, W.F.; Lee, S.L.; Havrda, M.C. Plasma-borne indicators of inflammasome activity in Parkinson′s disease patients. NPJ Park. Dis. 2021, 7, 2. [Google Scholar] [CrossRef]
  386. Wang, J.; Zhang, X.N.; Fang, J.N.; Hua, F.F.; Han, J.Y.; Yuan, Z.Q.; Xie, A.M. The mechanism behind activation of the Nod-like receptor family protein 3 inflammasome in Parkinson′s disease. Neural. Regen. Res. 2022, 17, 898–904. [Google Scholar] [CrossRef]
  387. Simola, N.; Morelli, M.; Carta, A.R. The 6-hydroxydopamine model of Parkinson′s disease. Neurotox. Res. 2007, 11, 151–167. [Google Scholar] [CrossRef]
  388. Si, X.L.; Fang, Y.J.; Li, L.F.; Gu, L.Y.; Yin, X.Z.; Tian, J.; Yan, Y.P.; Pu, J.L.; Zhang, B.R. From inflammasome to Parkinson′s disease: Does the NLRP3 inflammasome facilitate exosome secretion and exosomal alpha-synuclein transmission in Parkinson′s disease? Exp. Neurol. 2021, 336, 113525. [Google Scholar] [CrossRef]
  389. Chen, J.; Mao, K.; Yu, H.; Wen, Y.; She, H.; Zhang, H.; Liu, L.; Li, M.; Li, W.; Zou, F. p38-TFEB pathways promote microglia activation through inhibiting CMA-mediated NLRP3 degradation in Parkinson′s disease. J. Neuroinflamm. 2021, 18, 295. [Google Scholar] [CrossRef]
  390. Panicker, N.; Kam, T.I.; Wang, H.; Neifert, S.; Chou, S.C.; Kumar, M.; Brahmachari, S.; Jhaldiyal, A.; Hinkle, J.T.; Akkentli, F.; et al. Neuronal NLRP3 is a parkin substrate that drives neurodegeneration in Parkinson′s disease. Neuron 2022, 110, 2422–2437.e9. [Google Scholar] [CrossRef]
  391. Gris, D.; Ye, Z.; Iocca, H.A.; Wen, H.; Craven, R.R.; Gris, P.; Huang, M.; Schneider, M.; Miller, S.D.; Ting, J.P. NLRP3 plays a critical role in the development of experimental autoimmune encephalomyelitis by mediating Th1 and Th17 responses. J. Immunol. 2010, 185, 974–981. [Google Scholar] [CrossRef] [Green Version]
  392. Malhotra, S.; Río, J.; Urcelay, E.; Nurtdinov, R.; Bustamante, M.F.; Fernández, O.; Oliver, B.; Zettl, U.; Brassat, D.; Killestein, J.; et al. NLRP3 inflammasome is associated with the response to IFN-β in patients with multiple sclerosis. Brain 2015, 138, 644–652. [Google Scholar] [CrossRef] [Green Version]
  393. Malhotra, S.; Costa, C.; Eixarch, H.; Keller, C.W.; Amman, L.; Martínez-Banaclocha, H.; Midaglia, L.; Sarró, E.; Machín-Díaz, I.; Villar, L.M.; et al. NLRP3 inflammasome as prognostic factor and therapeutic target in primary progressive multiple sclerosis patients. Brain 2020, 143, 1414–1430. [Google Scholar] [CrossRef]
  394. Olcum, M.; Tastan, B.; Kiser, C.; Genc, S.; Genc, K. Microglial NLRP3 inflammasome activation in multiple sclerosis. Adv. Protein Chem. Struct. Biol. 2020, 119, 247–308. [Google Scholar] [CrossRef] [PubMed]
  395. Cui, Y.; Yu, H.; Bu, Z.; Wen, L.; Yan, L.; Feng, J. Focus on the Role of the NLRP3 Inflammasome in Multiple Sclerosis: Pathogenesis, Diagnosis, and Therapeutics. Front. Mol. Neurosci. 2022, 15, 894298. [Google Scholar] [CrossRef] [PubMed]
  396. Soares, J.L.; Oliveira, E.M.; Pontillo, A. Variants in NLRP3 and NLRC4 inflammasome associate with susceptibility and severity of multiple sclerosis. Mult. Scler. Relat. Disord. 2019, 29, 26–34. [Google Scholar] [CrossRef] [PubMed]
  397. Vidmar, L.; Maver, A.; Drulović, J.; Sepčić, J.; Novaković, I.; Ristič, S.; Šega, S.; Peterlin, B. Multiple Sclerosis patients carry an increased burden of exceedingly rare genetic variants in the inflammasome regulatory genes. Sci. Rep. 2019, 9, 9171. [Google Scholar] [CrossRef] [Green Version]
  398. Keane, R.W.; Dietrich, W.D.; de Rivero Vaccari, J.P. Inflammasome Proteins as Biomarkers of Multiple Sclerosis. Front. Neurol. 2018, 9, 135. [Google Scholar] [CrossRef] [PubMed]
  399. Kadowaki, A.; Quintana, F.J. The NLRP3 inflammasome in progressive multiple sclerosis. Brain 2020, 143, 1286–1288. [Google Scholar] [CrossRef] [PubMed]
  400. Farooqi, N.; Gran, B.; Constantinescu, C.S. Are current disease-modifying therapeutics in multiple sclerosis justified based on studies in experimental autoimmune encephalomyelitis? J. Neurochem. 2010, 115, 829–844. [Google Scholar] [CrossRef] [PubMed]
  401. Inoue, M.; Williams, K.L.; Gunn, M.D.; Shinohara, M.L. NLRP3 inflammasome induces chemotactic immune cell migration to the CNS in experimental autoimmune encephalomyelitis. Proc. Natl. Acad. Sci. USA 2012, 109, 10480–10485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Khan, N.; Kuo, A.; Brockman, D.A.; Cooper, M.A.; Smith, M.T. Pharmacological inhibition of the NLRP3 inflammasome as a potential target for multiple sclerosis induced central neuropathic pain. Inflammopharmacology 2018, 26, 77–86. [Google Scholar] [CrossRef] [PubMed]
  403. Bakhshi, S.; Shamsi, S. MCC950 in the treatment of NLRP3-mediated inflammatory diseases: Latest evidence and therapeutic outcomes. Int. Immunopharmacol. 2022, 106, 108595. [Google Scholar] [CrossRef]
  404. Hou, B.; Zhang, Y.; Liang, P.; He, Y.; Peng, B.; Liu, W.; Han, S.; Yin, J.; He, X. Inhibition of the NLRP3-inflammasome prevents cognitive deficits in experimental autoimmune encephalomyelitis mice via the alteration of astrocyte phenotype. Cell Death Dis. 2020, 11, 377. [Google Scholar] [CrossRef]
  405. Inoue, M.; Shinohara, M.L. The role of interferon-β in the treatment of multiple sclerosis and experimental autoimmune encephalomyelitis—In the perspective of inflammasomes. Immunology 2013, 139, 11–18. [Google Scholar] [CrossRef]
  406. Gros-Louis, F.; Gaspar, C.; Rouleau, G.A. Genetics of familial and sporadic amyotrophic lateral sclerosis. Biochim. Biophys. Acta 2006, 1762, 956–972. [Google Scholar] [CrossRef] [Green Version]
  407. Jaarsma, D.; Teuling, E.; Haasdijk, E.D.; De Zeeuw, C.I.; Hoogenraad, C.C. Neuron-specific expression of mutant superoxide dismutase is sufficient to induce amyotrophic lateral sclerosis in transgenic mice. J. Neurosci. 2008, 28, 2075–2088. [Google Scholar] [CrossRef] [Green Version]
  408. Yamanaka, K.; Boillee, S.; Roberts, E.A.; Garcia, M.L.; McAlonis-Downes, M.; Mikse, O.R.; Cleveland, D.W.; Goldstein, L.S. Mutant SOD1 in cell types other than motor neurons and oligodendrocytes accelerates onset of disease in ALS mice. Proc. Natl. Acad. Sci. USA 2008, 105, 7594–7599. [Google Scholar] [CrossRef] [Green Version]
  409. Mackenzie, I.R.; Rademakers, R.; Neumann, M. TDP-43 and FUS in amyotrophic lateral sclerosis and frontotemporal dementia. Lancet Neurol. 2010, 9, 995–1007. [Google Scholar] [CrossRef]
  410. Arai, T.; Hasegawa, M.; Akiyama, H.; Ikeda, K.; Nonaka, T.; Mori, H.; Mann, D.; Tsuchiya, K.; Yoshida, M.; Hashizume, Y.; et al. TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun. 2006, 351, 602–611. [Google Scholar] [CrossRef]
  411. Neumann, M.; Sampathu, D.M.; Kwong, L.K.; Truax, A.C.; Micsenyi, M.C.; Chou, T.T.; Bruce, J.; Schuck, T.; Grossman, M.; Clark, C.M.; et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science 2006, 314, 130–133. [Google Scholar] [CrossRef] [Green Version]
  412. Ezzi, S.A.; Urushitani, M.; Julien, J.P. Wild-type superoxide dismutase acquires binding and toxic properties of ALS-linked mutant forms through oxidation. J. Neurochem. 2007, 102, 170–178. [Google Scholar] [CrossRef]
  413. Roberts, K.; Zeineddine, R.; Corcoran, L.; Li, W.; Campbell, I.L.; Yerbury, J.J. Extracellular aggregated Cu/Zn superoxide dismutase activates microglia to give a cytotoxic phenotype. Glia 2013, 61, 409–419. [Google Scholar] [CrossRef]
  414. Zhao, W.; Beers, D.R.; Bell, S.; Wang, J.; Wen, S.; Baloh, R.H.; Appel, S.H. TDP-43 activates microglia through NF-κB and NLRP3 inflammasome. Exp. Neurol. 2015, 273, 24–35. [Google Scholar] [CrossRef] [PubMed]
  415. Banerjee, P.; Elliott, E.; Rifai, O.M.; O′Shaughnessy, J.; McDade, K.; Abrahams, S.; Chandran, S.; Smith, C.; Gregory, J.M. NLRP3 inflammasome as a key molecular target underlying cognitive resilience in amyotrophic lateral sclerosis. J. Pathol. 2022, 256, 262–268. [Google Scholar] [CrossRef] [PubMed]
  416. Kadhim, H.; Deltenre, P.; Martin, J.J.; Sébire, G. In-situ expression of Interleukin-18 and associated mediators in the human brain of sALS patients: Hypothesis for a role for immune-inflammatory mechanisms. Med. Hypotheses 2016, 86, 14–17. [Google Scholar] [CrossRef] [PubMed]
  417. Gugliandolo, A.; Giacoppo, S.; Bramanti, P.; Mazzon, E. NLRP3 Inflammasome Activation in a Transgenic Amyotrophic Lateral Sclerosis Model. Inflammation 2018, 41, 93–103. [Google Scholar] [CrossRef] [PubMed]
  418. Debye, B.; Schmülling, L.; Zhou, L.; Rune, G.; Beyer, C.; Johann, S. Neurodegeneration and NLRP3 inflammasome expression in the anterior thalamus of SOD1(G93A) ALS mice. Brain Pathol. 2018, 28, 14–27. [Google Scholar] [CrossRef]
  419. Michaelson, N.; Facciponte, D.; Bradley, W.; Stommel, E. Cytokine expression levels in ALS: A potential link between inflammation and BMAA-triggered protein misfolding. Cytokine Growth Factor Rev. 2017, 37, 81–88. [Google Scholar] [CrossRef]
  420. Van Schoor, E.; Ospitalieri, S.; Moonen, S.; Tomé, S.O.; Ronisz, A.; Ok, O.; Weishaupt, J.; Ludolph, A.C.; Van Damme, P.; Van Den Bosch, L.; et al. Increased pyroptosis activation in white matter microglia is associated with neuronal loss in ALS motor cortex. Acta Neuropathol. 2022, 144, 393–411. [Google Scholar] [CrossRef]
  421. Stallings, N.R.; Puttaparthi, K.; Luther, C.M.; Burns, D.K.; Elliott, J.L. Progressive motor weakness in transgenic mice expressing human TDP-43. Neurobiol. Dis. 2010, 40, 404–414. [Google Scholar] [CrossRef]
  422. Quek, H.; Cuní-López, C.; Stewart, R.; Colletti, T.; Notaro, A.; Nguyen, T.H.; Sun, Y.; Guo, C.C.; Lupton, M.K.; Roberts, T.L.; et al. ALS monocyte-derived microglia-like cells reveal cytoplasmic TDP-43 accumulation, DNA damage, and cell-specific impairment of phagocytosis associated with disease progression. J. Neuroinflamm. 2022, 19, 58. [Google Scholar] [CrossRef]
  423. Kenney, C.; Powell, S.; Jankovic, J. Autopsy-proven Huntington′s disease with 29 trinucleotide repeats. Mov. Disord. 2007, 22, 127–130. [Google Scholar] [CrossRef]
  424. Fusco, F.R.; Paldino, E. Role of Phosphodiesterases in Huntington′s Disease. Adv. Neurobiol. 2017, 17, 285–304. [Google Scholar] [CrossRef] [PubMed]
  425. Paldino, E.; Fusco, F.R. Emerging Role of NLRP3 Inflammasome/Pyroptosis in Huntington′s Disease. Int. J. Mol. Sci. 2022, 23, 8363. [Google Scholar] [CrossRef] [PubMed]
  426. Menalled, L.B.; Chesselet, M.F. Mouse models of Huntington′s disease. Trends Pharmacol. Sci. 2002, 23, 32–39. [Google Scholar] [CrossRef] [PubMed]
  427. Paldino, E.; D′Angelo, V.; Sancesario, G.; Fusco, F.R. Pyroptotic cell death in the R6/2 mouse model of Huntington′s disease: New insight on the inflammasome. Cell Death Discov. 2020, 6, 69. [Google Scholar] [CrossRef] [PubMed]
  428. Paldino, E.; D′Angelo, V.; Laurenti, D.; Angeloni, C.; Sancesario, G.; Fusco, F.R. Modulation of Inflammasome and Pyroptosis by Olaparib, a PARP-1 Inhibitor.; in the R6/2 Mouse Model of Huntington′s Disease. Cells 2020, 9, 2286. [Google Scholar] [CrossRef]
  429. Chen, K.P.; Hua, K.F.; Tsai, F.T.; Lin, T.Y.; Cheng, C.Y.; Yang, D.I.; Hsu, H.T.; Ju, T.C. A selective inhibitor of the NLRP3 inflammasome as a potential therapeutic approach for neuroprotection in a transgenic mouse model of Huntington′s disease. J. Neuroinflamm. 2022, 19, 56. [Google Scholar] [CrossRef]
  430. Siew, J.J.; Chen, H.M.; Chen, H.Y.; Chen, H.L.; Chen, C.M.; Soong, B.W.; Wu, Y.R.; Chang, C.P.; Chan, Y.C.; Lin, C.H.; et al. Galectin-3 is required for the microglia-mediated brain inflammation in a model of Huntington′s disease. Nat. Commun. 2019, 10, 3473. [Google Scholar] [CrossRef] [Green Version]
  431. Barake, F.; Soza, A.; González, A. Galectins in the brain: Advances in neuroinflammation, neuroprotection and therapeutic opportunities. Curr. Opin. Neurol. 2020, 33, 381–390. [Google Scholar] [CrossRef]
  432. Tricarico, P.M.; Caracciolo, I.; Crovella, S.; D′Agaro, P. Zika virus induces inflammasome activation in the glial cell line U87-MG. Biochem. Biophys. Res. Commun. 2017, 492, 597–602. [Google Scholar] [CrossRef]
  433. Zheng, Y.; Liu, Q.; Wu, Y.; Ma, L.; Zhang, Z.; Liu, T.; Jin, S.; She, Y.; Li, Y.P.; Cui, J. Zika virus elicits inflammation to evade antiviral response by cleaving cGAS via NS1-caspase-1 axis. EMBO J. 2018, 37, e99347. [Google Scholar] [CrossRef]
  434. Wang, W.; Li, G.; De, W.; Luo, Z.; Pan, P.; Tian, M.; Wang, Y.; Xiao, F.; Li, A.; Wu, K.; et al. Zika virus infection induces host inflammatory responses by facilitating NLRP3 inflammasome assembly and interleukin-1β secretion. Nat. Commun. 2018, 9, 106. [Google Scholar] [CrossRef] [Green Version]
  435. Ramos, H.J.; Lanteri, M.C.; Blahnik, G.; Negash, A.; Suthar, M.S.; Brassil, M.M.; Sodhi, K.; Treuting, P.M.; Busch, M.P.; Norris, P.J.; et al. IL-1β signaling promotes CNS-intrinsic immune control of West Nile virus infection. PLoS Pathog. 2012, 8, e1003039. [Google Scholar] [CrossRef] [Green Version]
  436. Yang, C.M.; Lin, C.C.; Lee, I.T.; Lin, Y.H.; Yang, C.M.; Chen, W.J.; Jou, M.J.; Hsiao, L.D. Japanese encephalitis virus induces matrix metalloproteinase-9 expression via a ROS/c-Src/PDGFR/PI3K/Akt/MAPKs-dependent AP-1 pathway in rat brain astrocytes. J. Neuroinflamm. 2012, 9, 12. [Google Scholar] [CrossRef] [Green Version]
  437. He, W.; Zhao, Z.; Anees, A.; Li, Y.; Ashraf, U.; Chen, Z.; Song, Y.; Chen, H.; Cao, S.; Ye, J. p21-activated kinase 4 signaling promotes Japanese encephalitis virus-mediated inflammation in astrocytes. Front. Cell. Infect. Microbiol. 2017, 7, 271. [Google Scholar] [CrossRef]
  438. Ashraf, U.; Ding, Z.; Deng, S.; Ye, J.; Cao, S.; Chen, Z. Pathogenicity and virulence of Japanese encephalitis virus: Neuroinflammation and neuronal cell damage. Virulence 2021, 12, 968–980. [Google Scholar] [CrossRef]
  439. Burdo, T.H.; Lackner, A.; Williams, K.C. Monocyte/macrophages, and their role in HIV neuropathogenesis. Immunol. Rev. 2013, 254, 102–113. [Google Scholar] [CrossRef] [Green Version]
  440. Walsh, J.G.; Reinke, S.N.; Mamik, M.K.; McKenzie, B.A.; Maingat, F.; Branton, W.G.; Broadhurst, D.I.; Power, C. Rapid inflammasome activation in microglia contributes to brain disease in HIV/AIDS. Retrovirology 2014, 11, 35. [Google Scholar] [CrossRef] [Green Version]
  441. Breitinger, U.; Farag, N.S.; Sticht, H.; Breitinger, H.G. Viroporins: Structure, function, and their role in the life cycle of SARS-CoV-2. Int. J. Biochem. Cell Biol. 2022, 145, 106185. [Google Scholar] [CrossRef]
  442. Poeck, H.; Bscheider, M.; Gross, O.; Finger, K.; Roth, S.; Rebsamen, M.; Hannesschläger, N.; Schlee, M.; Rothenfusser, S.; Barchet, W.; et al. Recognition of RNA virus by RIG-I results in activation of CARD9 and inflammasome signaling for interleukin 1 beta production. Nat. Immunol. 2010, 11, 63–69, Addendum in Nat. Immunol. 2014, 15, 109. [Google Scholar] [CrossRef]
  443. Rajan, J.V.; Rodriguez, D.; Miao, E.A.; Aderem, A. The NLRP3 inflammasome detects encephalomyocarditis virus and vesicular stomatitis virus infection. J. Virol. 2011, 85, 4167–4172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  444. Szabo, M.P.; Iba, M.; Nath, A.; Masliah, E.; Kim, C. Does SARS-CoV-2 affect neurodegenerative disorders? TLR2, a potential receptor for SARS-CoV-2 in the CNS. Exp. Mol. Med. 2022, 54, 447–454. [Google Scholar] [CrossRef] [PubMed]
  445. Hung, E.C.; Chim, S.S.; Chan, P.K.; Tong, Y.K.; Ng, E.K.; Chiu, R.W.; Leung, C.B.; Sung, J.J.; Tam, J.S.; Lo, Y.M. Detection of SARS coronavirus RNA in the cerebrospinal fluid of a patient with severe acute respiratory syndrome. Clin. Chem. 2003, 49, 2108–2109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  446. Sepehrinezhad, A.; Rezaeitalab, F.; Shahbazi, A.; Sahab-Negah, S. A Computational-Based Drug Repurposing Method Targeting SARS-CoV-2 and its Neurological Manifestations Genes and Signaling Pathways. Bioinform. Biol. Insights. 2021, 15, 11779322211026728. [Google Scholar] [CrossRef] [PubMed]
  447. Chen, R.; Wang, K.; Yu, J.; Howard, D.; French, L.; Chen, Z.; Wen, C.; Xu, Z. The spatial and cell-type distribution of SARS-CoV-2 receptor ACE2 in the human and mouse brains. Front. Neurol. 2020, 11, 573095. [Google Scholar] [CrossRef] [PubMed]
  448. Ribeiro, D.E.; Oliveira-Giacomelli, A.; Glaser, T.; Arnaud-Sampaio, V.F.; Andrejew, R.; Dieckmann, L.; Baranova, J.; Lameu, C.; Ratajczak, M.Z.; Ulrich, H. Hyperactivation of P2X7 receptors as a culprit of COVID-19 neuropathology. Mol. Psychiatry 2021, 26, 1044–1059. [Google Scholar] [CrossRef]
  449. Sepehrinezhad, A.; Gorji, A.; Sahab Negah, S. SARS-CoV-2 may trigger inflammasome and pyroptosis in the central nervous system: A mechanistic view of neurotropism. Inflammopharmacology 2021, 29, 1049–1059. [Google Scholar] [CrossRef]
  450. Helms, J.; Kremer, S.; Merdji, H.; Clere-Jehl, R.; Schenck, M.; Kummerlen, C.; Collange, O.; Boulay, C.; Fafi-Kremer, S.; Ohana, M.; et al. Neurologic features in severe SARS-CoV-2 infection. N. Engl. J. Med. 2020, 382, 2268–2270. [Google Scholar] [CrossRef]
  451. Zhao, Y.; Li, W.; Lukiw, W. Ubiquity of the SARS-CoV-2 receptor ACE2 and upregulation in limbic regions of Alzheimer′s disease brain. Folia Neuropathol. 2021, 59, 232–238. [Google Scholar] [CrossRef]
  452. Ding, Q.; Shults, N.V.; Gychka, S.G.; Harris, B.T.; Suzuki, Y.J. Protein Expression of Angiotensin-Converting Enzyme 2 (ACE2) is Upregulated in Brains with Alzheimer′s Disease. Int. J. Mol. Sci. 2021, 22, 1687. [Google Scholar] [CrossRef]
  453. Theobald, S.J.; Simonis, A.; Georgomanolis, T.; Kreer, C.; Zehner, M.; Eisfeld, H.S.; Albert, M.C.; Chhen, J.; Motameny, S.; Erger, F.; et al. Long-lived macrophage reprogramming drives spike protein-mediated inflammasome activation in COVID-19. EMBO Mol. Med. 2021, 13, e14150. [Google Scholar] [CrossRef]
  454. Xu, H.; Akinyemi, I.A.; Chitre, S.A.; Loeb, J.C.; Lednicky, J.A.; McIntosh, M.T.; Bhaduri-McIntosh, S. SARS-CoV-2 viroporin encoded by ORF3a triggers the NLRP3 inflammatory pathway. Virology 2022, 568, 13–22. [Google Scholar] [CrossRef]
  455. Freeman, T.L.; Swartz, T.H. Targeting the NLRP3 Inflammasome in Severe COVID-19. Front. Immunol. 2020, 11, 1518. [Google Scholar] [CrossRef]
  456. Fatima, S.; Zaidi, S.S.; Alsharidah, A.S.; Aljaser, F.S.; Banu, N. Possible Prophylactic Approach for SARS-CoV-2 Infection by Combination of Melatonin, Vitamin C and Zinc in Animals. Front. Vet. Sci. 2020, 7, 585789. [Google Scholar] [CrossRef]
  457. Ding, H.G.; Deng, Y.Y.; Yang, R.Q.; Wang, Q.S.; Jiang, W.Q.; Han, Y.L.; Huang, L.Q.; Wen, M.Y.; Zhong, W.H.; Li, X.S.; et al. Hypercapnia induces IL-1β overproduction via activation of NLRP3 inflammasome: Implication in cognitive impairment in hypoxemic adult rats. J. Neuroinflamm. 2018, 15, 4. [Google Scholar] [CrossRef] [Green Version]
  458. Heneka, M.T.; Golenbock, D.; Latz, E.; Morgan, D.; Brown, R. Immediate and long-term consequences of COVID-19 infections for the development of neurological disease. Alzheimers Res. Ther. 2020, 12, 69. [Google Scholar] [CrossRef]
  459. Farheen, S.; Agrawal, S.; Zubair, S.; Agrawal, A.; Jamal, F.; Altaf, I.; Kashif Anwar, A.; Umair, S.M.; Owais, M. Pathophysiology of aging and immune-senescence: Possible correlates with comorbidity and mortality in middle-aged and old COVID-19 patients. Front. Aging 2021, 2, 748591. [Google Scholar] [CrossRef]
  460. Flud, V.V.; Shcherbuk, Y.A.; Shcherbuk, A.Y.; Leonov, V.I.; Al-Sahli, O.A. Neurological complications and consequences of new coronavirus COVID-19 infection in elderly and old patients (literature review). Adv. Gerontol. 2022, 35, 231–242. [Google Scholar]
  461. Fu, Y.W.; Xu, H.S.; Liu, S.J. COVID-19 and neurodegenerative diseases. Eur. Rev. Med. Pharmacol. Sci. 2022, 26, 4535–4544. [Google Scholar] [CrossRef]
  462. Baazaoui, N.; Iqbal, K. COVID-19 and Neurodegenerative Diseases: Prion-Like Spread and Long-Term Consequences. J. Alzheimers Dis. 2022, 88, 399–416. [Google Scholar] [CrossRef]
  463. Bernardini, A.; Gigli, G.L.; Janes, F.; Pellitteri, G.; Ciardi, C.; Fabris, M.; Valente, M. Creutzfeldt-Jakob disease after COVID-19: Infection-induced prion protein misfolding? A case report. Prion 2022, 16, 78–83. [Google Scholar] [CrossRef]
  464. Tetz, G.; Tetz, V. Prion-like Domains in Spike Protein of SARS-CoV-2 Differ across Its Variants and Enable Changes in Affinity to ACE2. Microorganisms 2022, 10, 280. [Google Scholar] [CrossRef] [PubMed]
  465. Wilson, L.; Stewart, W.; Dams-O′Connor, K.; Diaz-Arrastia, R.; Horton, L.; Menon, D.K.; Polinder, S. The chronic and evolving neurological consequences of traumatic brain injury. Lancet Neurol. 2017, 16, 813–825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  466. Lee, J.H.; Kim, H.J.; Kim, J.U.; Yook, T.H.; Kim, K.H.; Lee, J.Y.; Yang, G. A Novel Treatment Strategy by Natural Products in NLRP3 Inflammasome-Mediated Neuroinflammation in Alzheimer′s and Parkinson′s Disease. Int. J. Mol. Sci. 2021, 22, 1324. [Google Scholar] [CrossRef] [PubMed]
  467. Wu, A.G.; Zhou, X.G.; Qiao, G.; Yu, L.; Tang, Y.; Yan, L.; Qiu, W.Q.; Pan, R.; Yu, C.L.; Law, B.Y.; et al. Targeting microglial autophagic degradation in NLRP3 inflammasome-mediated neurodegenerative diseases. Ageing Res. Rev. 2021, 65, 101202. [Google Scholar] [CrossRef]
  468. Freeman, L.; Guo, H.; David, C.N.; Brickey, W.J.; Jha, S.; Ting, J.P. NLR members NLRC4 and NLRP3 mediate sterile inflammasome activation in microglia and astrocytes. J. Exp. Med. 2017, 214, 1351–1370. [Google Scholar] [CrossRef] [Green Version]
  469. McKenzie, B.A.; Mamik, M.K.; Saito, L.B.; Boghozian, R.; Monaco, M.C.; Major, E.O.; Lu, J.Q.; Branton, W.G.; Power, C. Caspase-1 inhibition prevents glial inflammasome activation and pyroptosis in models of multiple sclerosis. Proc. Natl. Acad. Sci. USA 2018, 115, E6065–E6074. [Google Scholar] [CrossRef] [Green Version]
  470. Yap, J.K.Y.; Pickard, B.S.; Chan, E.W.L.; Gan, S.Y. The Role of Neuronal NLRP1 Inflammasome in Alzheimer′s Disease: Bringing Neurons into the Neuroinflammation Game. Mol. Neurobiol. 2019, 56, 7741–7753. [Google Scholar] [CrossRef]
  471. Nagyőszi, P.; Nyúl-Tóth, Á.; Fazakas, C.; Wilhelm, I.; Kozma, M.; Molnár, J.; Haskó, J.; Krizbai, I.A. Regulation of NOD-like receptors and inflammasome activation in cerebral endothelial cells. J. Neurochem. 2015, 135, 551–564. [Google Scholar] [CrossRef]
  472. Feng, Y.S.; Tan, Z.X.; Wu, L.Y.; Dong, F.; Zhang, F. The involvement of NLRP3 inflammasome in the treatment of neurodegenerative diseases. Biomed. Pharmacother. 2021, 38, 111428. [Google Scholar] [CrossRef]
  473. Lahooti, B.; Chhibber, T.; Bagchi, S.; Varahachalam, S.P.; Jayant, R.D. Therapeutic role of inflammasome inhibitors in neurodegenerative disorders. Brain Behav. Immun. 2021, 91, 771–783. [Google Scholar] [CrossRef]
  474. Corcoran, S.E.; Halai, R.; Cooper, M.A. Pharmacological Inhibition of the Nod-Like Receptor Family Pyrin Domain Containing 3 Inflammasome with MCC950. Pharmacol. Rev. 2021, 73, 968–1000. [Google Scholar] [CrossRef]
  475. Soriano-Teruel, P.M.; García-Laínez, G.; Marco-Salvador, M.; Pardo, J.; Arias, M.; DeFord, C.; Merfort, I.; Vicent, M.J.; Pelegrín, P.; Sancho, M.; et al. Identification of an ASC oligomerization inhibitor for the treatment of inflammatory diseases. Cell Death Dis. 2021, 12, 1155. [Google Scholar] [CrossRef]
  476. De Souza, N. Model organisms: Mouse models challenged. Nat. Methods 2013, 10, 288. [Google Scholar] [CrossRef]
  477. Pound, P.; Ritskes-Hoitinga, M. Is it possible to overcome issues of external validity in preclinical animal research? Why most animal models are bound to fail. J. Transl. Med. 2018, 16, 304. [Google Scholar] [CrossRef] [Green Version]
  478. Gharib, W.H.; Robinson-Rechavi, M. When orthologs diverge between human and mouse. Brief. Bioinform. 2011, 12, 436–441. [Google Scholar] [CrossRef] [Green Version]
  479. Seok, J.; Warren, H.S.; Cuenca, A.G.; Mindrinos, M.N.; Baker, H.V.; Xu, W.; Richards, D.R.; Mcdonald-Smith, G.P.; Gao, H.; Hennessy, L.; et al. Inflammation and Host Response to Injury, Large Scale Collaborative Research Program. Genomic responses in mouse models poorly mimic human inflammatory diseases. Proc. Natl. Acad. Sci. USA 2013, 110, 3507–3512. [Google Scholar] [CrossRef] [Green Version]
  480. Greek, R.; Menache, A. Systematic reviews of animal models: Methodology versus epistemology. Int. J. Med. Sci. 2013, 10, 206–221. [Google Scholar] [CrossRef] [Green Version]
  481. Akhtar, A. The flaws and human harms of animal experimentation. Camb. Q. Health Ethics. 2015, 24, 407–419. [Google Scholar] [CrossRef] [Green Version]
  482. Balls, M. It′s Time to Include Harm to Humans in Harm-Benefit Analysis—But How to Do It, that is the Question. Altern. Lab. Anim. 2021, 49, 182–196. [Google Scholar] [CrossRef]
  483. Chiarini, A.; Gardenal, E.; Whitfield, J.F.; Chakravarthy, B.; Armato, U.; Dal Pra, I. Preventing the spread of Alzheimer′s disease neuropathology: A role for calcilytics? Curr. Pharm. Biotechnol. 2015, 16, 696–706. [Google Scholar] [CrossRef]
  484. Chiarini, A.; Armato, U.; Liu, D.; Dal Prà, I. Calcium-Sensing Receptor Antagonist NPS 2143 Restores Amyloid Precursor Protein Physiological Non-Amyloidogenic Processing in Aβ-Exposed Adult Human Astrocytes. Sci. Rep. 2017, 7, 1277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  485. Van Dyck, C.H.; Swanson, C.J.; Aisen, P.; Bateman, R.J.; Chen, C.; Gee, M.; Kanekiyo, M.; Li, D.; Reyderman, L.; Cohen, S.; et al. Lecanemab in Early Alzheimer′s Disease. N. Engl. J. Med. 2023, 388, 9–21. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of stressors and factors inducing/modulating NLRP3 inflammasome’s activation and its sequels in astrocytes and microglia under acute injuries due to hypoxic ischemia, stroke, and hemorrhage. Left: Astrocyte’s prompt response. Acute O2 tension fall activates Ca2+ influx through TPRV1 channels, triggering the JAK2/STAT3 axis and NLRP3 inflammasome activation. It also increases BACE1 and IL-33 gene expression. Over-released IL-33 binds its STD2 receptor, whose signaling mitigates NLRP3 activity. Later, BACE1 increased activity overproduces Aβs. Extracellularly released excess Aβs bind and activate CaSR signaling, which contributes to NLRP3 inflammasome activation by reducing cAMP levels and activating CaMKII. Aβ•CaSR signaling also increases BACE1 and GSK-3β activities, driving the over production of Aβs from APP and p-Taues, which are both intracellularly accumulated and extracellularly released. CaSR NAM (Calcilytic) NPS2143 and CaMKII inhibitor KN93 suppress Aβs•CaSR signaling noxious effects (see for more details Box 1). Top right: Late wild-type microglia response. The NLRP3 activation is blocked by various agents, which activate via Akt the expression of NRF2 transcription factor. NRF2 activity reduces the M1 (proinflammatory) fraction of microglia. Bottom right: In a model of NLRP3 full-knockout microglia Ca2+ influx activates in NLRP3 stead the AIM2 inflammasome’s signaling, the upshot being the same, i.e., the overproduction/release of IL-1β and IL-18 [336]. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange colored. Abbreviations: Aβs = amyloid-β peptides; AC = adenylyl cyclase; AIM2 = absent in melanoma 2 inflammasome; Akt = protein kinase B; APP = amyloid precursor protein; ASC = apoptosis-associated speck-like protein endowed with a caspase recruitment domain or CARD; BACE1 = β-secretase; BBB = blood-brain barrier; cAMP = 3′,5′-cyclic adenosine monophosphate; CASP1 = caspase-1; CaMKII = Ca2+/calmodulin-dependent protein kinase II; CaSR, calcium-sensing receptor; GdCl3 = gadolinium chloride; GSK-3β = glycogen synthase kinase-3β; JAK2 = Janus kinase 2; KN93 = N-[2-[[[(E)-3-(4-chlorophenyl)prop-2-enyl]-methylamino]methyl]phenyl]-N-(2-hydroxyethyl)-4-methoxybenzenesulfon-amide; NPS-2143 = 2-chloro-6-[(2R)-2-hydroxy-3-[(2-methyl-1-naphthalen-2-ylpropan-2-yl)amino]-propoxy]-benzonitrile; p-Taues = hyperphosphorylated Tau proteins; STAT3 = signal transducer and activator of transcription 3); STD2 = suppression of tumorigenicity 2 (receptor); TPRV1 = vanilloid type 1 receptor/channel; WT = wild-type. ↓O2 = decrease in oxygen tension. The other arrows show the sequences of molecular events induced by stressors and factors. ⊥ = inhibition.
Figure 1. Schematic illustration of stressors and factors inducing/modulating NLRP3 inflammasome’s activation and its sequels in astrocytes and microglia under acute injuries due to hypoxic ischemia, stroke, and hemorrhage. Left: Astrocyte’s prompt response. Acute O2 tension fall activates Ca2+ influx through TPRV1 channels, triggering the JAK2/STAT3 axis and NLRP3 inflammasome activation. It also increases BACE1 and IL-33 gene expression. Over-released IL-33 binds its STD2 receptor, whose signaling mitigates NLRP3 activity. Later, BACE1 increased activity overproduces Aβs. Extracellularly released excess Aβs bind and activate CaSR signaling, which contributes to NLRP3 inflammasome activation by reducing cAMP levels and activating CaMKII. Aβ•CaSR signaling also increases BACE1 and GSK-3β activities, driving the over production of Aβs from APP and p-Taues, which are both intracellularly accumulated and extracellularly released. CaSR NAM (Calcilytic) NPS2143 and CaMKII inhibitor KN93 suppress Aβs•CaSR signaling noxious effects (see for more details Box 1). Top right: Late wild-type microglia response. The NLRP3 activation is blocked by various agents, which activate via Akt the expression of NRF2 transcription factor. NRF2 activity reduces the M1 (proinflammatory) fraction of microglia. Bottom right: In a model of NLRP3 full-knockout microglia Ca2+ influx activates in NLRP3 stead the AIM2 inflammasome’s signaling, the upshot being the same, i.e., the overproduction/release of IL-1β and IL-18 [336]. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange colored. Abbreviations: Aβs = amyloid-β peptides; AC = adenylyl cyclase; AIM2 = absent in melanoma 2 inflammasome; Akt = protein kinase B; APP = amyloid precursor protein; ASC = apoptosis-associated speck-like protein endowed with a caspase recruitment domain or CARD; BACE1 = β-secretase; BBB = blood-brain barrier; cAMP = 3′,5′-cyclic adenosine monophosphate; CASP1 = caspase-1; CaMKII = Ca2+/calmodulin-dependent protein kinase II; CaSR, calcium-sensing receptor; GdCl3 = gadolinium chloride; GSK-3β = glycogen synthase kinase-3β; JAK2 = Janus kinase 2; KN93 = N-[2-[[[(E)-3-(4-chlorophenyl)prop-2-enyl]-methylamino]methyl]phenyl]-N-(2-hydroxyethyl)-4-methoxybenzenesulfon-amide; NPS-2143 = 2-chloro-6-[(2R)-2-hydroxy-3-[(2-methyl-1-naphthalen-2-ylpropan-2-yl)amino]-propoxy]-benzonitrile; p-Taues = hyperphosphorylated Tau proteins; STAT3 = signal transducer and activator of transcription 3); STD2 = suppression of tumorigenicity 2 (receptor); TPRV1 = vanilloid type 1 receptor/channel; WT = wild-type. ↓O2 = decrease in oxygen tension. The other arrows show the sequences of molecular events induced by stressors and factors. ⊥ = inhibition.
Biomedicines 11 00999 g001
Figure 2. Schematic depiction of stressors and factors inducing/modulating glial cell NLRP3 inflammasome activation and its consequences in AD. Exogenous Aβs, p-Taues, ATP, ASC, IL-1β, and IL-18 interact with cell surface receptors, including CaSR (see Box 1), TL-4, and P2X7 (see Box 2), or are endocytosed to activate NF-κB and NLRP3 inflammasome signaling. They also induce ER stress, release Cathepsin B from damaged lysosomes, and block autophagy, while over-releasing further amounts of Aβs, p-Taues, and inflammatory cytokines. Altogether, they damage myelin sheaths and cause M1 microglial phenotype polarization and neuron and oligodendrocyte pyroptotic death. NLRP3 and receptor inhibitors mitigate the just-mentioned noxious effects. Additionally, the CaSR NAM NPS-2143 blocks Aβs, p-Taues, and IL-6 over production and release and reactivates autophagy (not shown; [181,199,354]). Regarding the roles of other-than-NLRP3 inflammasomes, see [24]. A yellow frame encloses the assembled NLRP3 inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: A438079 = 3-[[5-(2,3-dichlorophenyl)tetrazol-1-yl]methyl]pyridine; Aβs = amyloid-β peptides; ASC = apoptosis-associated speck-like protein endowed with a CARD; Bay117082 = (E)-3-(4-methylphenyl)sulfonylprop-2-enenitrile; BBB = blood-brain barrier; BBG = brilliant blue G; cAMP, 3′,5′-cyclic adenosine monophosphate; CA074 = CAS 134448-10-5; CASP1 = caspase-1; CaSR = calcium-sensing receptor; CCL3 = gene encoding MIP-1α chemokine; DHM = dihydromyricetin; E2 = estradiol; FOXO1 = forkhead box protein O1; GMF, glia maturation factor; JAK2 = Janus kinase 2; Keap1 = Kelch-like ECH-associated protein 1; KN93 = N-[2-[[[(E)-3-(4-chlorophenyl)prop-2-enyl]-methylamino]methyl]-phenyl]-N-(2-hydroxyethyl)-4-methoxybenzenesulfon-amide; LPS = bacterial lipopolysaccharide; MCC950, CAS 210826-40-7; MIP-1α = monocyte chemoattractant protein-1α; mTOR = mammalian target of rapamycin; MyD88 = myeloid differentiation primary response 88; NF-κB = nuclear factor κB; P2X7 = purinergic receptor; p-Taues = hyperphosphorylated Tau proteins; STAT3 = signal transducer and activator of transcription 3; TLR-4 = Toll-like receptor 4; TPRV1, vanilloid type 1 receptor/channel; TXNIP = thioredoxin interacting protein. The small arrows close to a name indicate (↓) decrease, or (↑) increase in levels. ⊥ = inhibition. The other arrows show the sequences of molecular events induced by stressors and factors.
Figure 2. Schematic depiction of stressors and factors inducing/modulating glial cell NLRP3 inflammasome activation and its consequences in AD. Exogenous Aβs, p-Taues, ATP, ASC, IL-1β, and IL-18 interact with cell surface receptors, including CaSR (see Box 1), TL-4, and P2X7 (see Box 2), or are endocytosed to activate NF-κB and NLRP3 inflammasome signaling. They also induce ER stress, release Cathepsin B from damaged lysosomes, and block autophagy, while over-releasing further amounts of Aβs, p-Taues, and inflammatory cytokines. Altogether, they damage myelin sheaths and cause M1 microglial phenotype polarization and neuron and oligodendrocyte pyroptotic death. NLRP3 and receptor inhibitors mitigate the just-mentioned noxious effects. Additionally, the CaSR NAM NPS-2143 blocks Aβs, p-Taues, and IL-6 over production and release and reactivates autophagy (not shown; [181,199,354]). Regarding the roles of other-than-NLRP3 inflammasomes, see [24]. A yellow frame encloses the assembled NLRP3 inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: A438079 = 3-[[5-(2,3-dichlorophenyl)tetrazol-1-yl]methyl]pyridine; Aβs = amyloid-β peptides; ASC = apoptosis-associated speck-like protein endowed with a CARD; Bay117082 = (E)-3-(4-methylphenyl)sulfonylprop-2-enenitrile; BBB = blood-brain barrier; BBG = brilliant blue G; cAMP, 3′,5′-cyclic adenosine monophosphate; CA074 = CAS 134448-10-5; CASP1 = caspase-1; CaSR = calcium-sensing receptor; CCL3 = gene encoding MIP-1α chemokine; DHM = dihydromyricetin; E2 = estradiol; FOXO1 = forkhead box protein O1; GMF, glia maturation factor; JAK2 = Janus kinase 2; Keap1 = Kelch-like ECH-associated protein 1; KN93 = N-[2-[[[(E)-3-(4-chlorophenyl)prop-2-enyl]-methylamino]methyl]-phenyl]-N-(2-hydroxyethyl)-4-methoxybenzenesulfon-amide; LPS = bacterial lipopolysaccharide; MCC950, CAS 210826-40-7; MIP-1α = monocyte chemoattractant protein-1α; mTOR = mammalian target of rapamycin; MyD88 = myeloid differentiation primary response 88; NF-κB = nuclear factor κB; P2X7 = purinergic receptor; p-Taues = hyperphosphorylated Tau proteins; STAT3 = signal transducer and activator of transcription 3; TLR-4 = Toll-like receptor 4; TPRV1, vanilloid type 1 receptor/channel; TXNIP = thioredoxin interacting protein. The small arrows close to a name indicate (↓) decrease, or (↑) increase in levels. ⊥ = inhibition. The other arrows show the sequences of molecular events induced by stressors and factors.
Biomedicines 11 00999 g002
Figure 3. Summary illustration of stressors and factors inducing/modulating dopaminergic neurons’ and microglia’s NLRP3 inflammasome activation and its consequences in PD. Overproduced α-synuclein (α-Syn) forms cytosolic aggregates (named when massive Lewis bodies) that damage lysosomes releasing cathepsin B, a cysteine protease. The latter interferes with mitochondrial activities causing in sequence dysfunctional mitophagy, ROS surpluses, oxidative stress, NF-κB pathway signaling, and overexpression of NLRP3 inflammasome components, the latter’s activation, and its downstream consequences. Exogenous ATP from pyroptotic cells helps activate NLRP3 inflammasome via the P2X7 purinergic receptor signaling (see Box 2 for more details). The upshots are the release of IL-1β and IL-18 and K+ efflux through pores made of GSDMD-N terminal fragments. α-Syn is also released extracellularly within exosomes that spread and are taken up by neighboring neural cells, expanding the neuropathology, or they circulate in the body fluids thus affecting peripheral tissues. Accumulated Cu2+ ions also harm mitochondria contributing to NLRP3 inflammasome’s activation. The toxic α-Syn effects are similar in microglia, in which they are mediated by TLR-2 and TLR-4 receptors too. α-Syn also blocks the chaperone-mediated autophagy (CMA) pathway regulated by the p38 MAPK/TEFB axis. Eventually, both nigrostriatal dopaminergic neurons and microglia undergo pyroptotic death. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: ASC = apoptosis-associated speck-like protein endowed with a CARD domain; 5-BDBD = 5-(3-Bromophenyl)-1,3-dihydro-2H-benzofuro[3,2-e]-1,4-diazepin-2-one; CASP1 = caspase-1; Exos = exosomes; GSDMD-N = gasdermin D N-terminal fragments; NF-κB = nuclear factor κB; P2X7 = purinergic receptor; p38 MAPK = p38 mitogen activated protein kinase; ROS = reactive oxygen species; TFEB = transcription factor EB; TLR=Toll-like receptor. ↑ROS = increase in ROS levels. ⊥ = inhibition. The other arrows show the sequences of molecular events induced by stressors and factors.
Figure 3. Summary illustration of stressors and factors inducing/modulating dopaminergic neurons’ and microglia’s NLRP3 inflammasome activation and its consequences in PD. Overproduced α-synuclein (α-Syn) forms cytosolic aggregates (named when massive Lewis bodies) that damage lysosomes releasing cathepsin B, a cysteine protease. The latter interferes with mitochondrial activities causing in sequence dysfunctional mitophagy, ROS surpluses, oxidative stress, NF-κB pathway signaling, and overexpression of NLRP3 inflammasome components, the latter’s activation, and its downstream consequences. Exogenous ATP from pyroptotic cells helps activate NLRP3 inflammasome via the P2X7 purinergic receptor signaling (see Box 2 for more details). The upshots are the release of IL-1β and IL-18 and K+ efflux through pores made of GSDMD-N terminal fragments. α-Syn is also released extracellularly within exosomes that spread and are taken up by neighboring neural cells, expanding the neuropathology, or they circulate in the body fluids thus affecting peripheral tissues. Accumulated Cu2+ ions also harm mitochondria contributing to NLRP3 inflammasome’s activation. The toxic α-Syn effects are similar in microglia, in which they are mediated by TLR-2 and TLR-4 receptors too. α-Syn also blocks the chaperone-mediated autophagy (CMA) pathway regulated by the p38 MAPK/TEFB axis. Eventually, both nigrostriatal dopaminergic neurons and microglia undergo pyroptotic death. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: ASC = apoptosis-associated speck-like protein endowed with a CARD domain; 5-BDBD = 5-(3-Bromophenyl)-1,3-dihydro-2H-benzofuro[3,2-e]-1,4-diazepin-2-one; CASP1 = caspase-1; Exos = exosomes; GSDMD-N = gasdermin D N-terminal fragments; NF-κB = nuclear factor κB; P2X7 = purinergic receptor; p38 MAPK = p38 mitogen activated protein kinase; ROS = reactive oxygen species; TFEB = transcription factor EB; TLR=Toll-like receptor. ↑ROS = increase in ROS levels. ⊥ = inhibition. The other arrows show the sequences of molecular events induced by stressors and factors.
Biomedicines 11 00999 g003
Figure 4. Summary depiction of stressors and factors inducing/modulating motoneurons’ and microglia’s NLRP3 inflammasome’s activation and its consequences in ALS. Mutated/misfolded SOD1 and TDP-43 proteins as variously sized aggregates damage mitochondria, causing in sequence ROS surpluses release, oxidative stress, and NF-κB pathway signaling. These lead to NLRP3 inflammasome’s component overexpression, NLRP3 inflammasome activation, over-release of IL-1β, IL-18, K+ efflux, and eventually motoneurons’ and microglia’s pyroptosis. Exposure to the toxic BMMA amino-acid released by Cyanobacteria worsens the toxic effects of misfolded/mutated SOD1 and TDP-43. Toll-like receptors and CD-14 bind misfolded/mutated SOD1 and TDP-43 activating the AP1/NF-κB axis, and the expression and activation of NLRP3 inflammasome’s components. ATP from pyroptotic cells partakes in NLRP3 activation via P2X7 receptor signaling (see Box 2 for details). Astrocytes also release misfolded/mutated SOD1 and TDP3 that are engulfed by other neural cells, thus spreading the neuropathology. Besides ATP, pyroptotic cells also release NLRP3, SOD1, TDP-43, and ASC proteins that contribute to the neuroinflammation. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: AP1 = activator protein 1; ASC = apoptosis-associated speck-like protein endowed with a CARD domain; BMAA = β-methylamino-L-alanine; CASP1 = caspase-1; CD14 = cluster of differentiation 14; GSDMD-N = gasdermin D N-terminal fragments; NF-κB = nuclear factor κB; NOX2 = NADPH oxidase 2; P2X7R = purinergic receptor; ROS = reactive oxygen species; SOD1 = superoxide dismutase 1; TDP-43 = TAR DNA-binding protein 43; TLR− =Toll-like receptor−; TNF-α = tumor necrosis factor-α; WT = wild-type. The arrows show the sequences of molecular events induced by stressors and factors.
Figure 4. Summary depiction of stressors and factors inducing/modulating motoneurons’ and microglia’s NLRP3 inflammasome’s activation and its consequences in ALS. Mutated/misfolded SOD1 and TDP-43 proteins as variously sized aggregates damage mitochondria, causing in sequence ROS surpluses release, oxidative stress, and NF-κB pathway signaling. These lead to NLRP3 inflammasome’s component overexpression, NLRP3 inflammasome activation, over-release of IL-1β, IL-18, K+ efflux, and eventually motoneurons’ and microglia’s pyroptosis. Exposure to the toxic BMMA amino-acid released by Cyanobacteria worsens the toxic effects of misfolded/mutated SOD1 and TDP-43. Toll-like receptors and CD-14 bind misfolded/mutated SOD1 and TDP-43 activating the AP1/NF-κB axis, and the expression and activation of NLRP3 inflammasome’s components. ATP from pyroptotic cells partakes in NLRP3 activation via P2X7 receptor signaling (see Box 2 for details). Astrocytes also release misfolded/mutated SOD1 and TDP3 that are engulfed by other neural cells, thus spreading the neuropathology. Besides ATP, pyroptotic cells also release NLRP3, SOD1, TDP-43, and ASC proteins that contribute to the neuroinflammation. A yellow frame encloses the assembled inflammasomes, while nuclear envelopes are orange-colored. Abbreviations: AP1 = activator protein 1; ASC = apoptosis-associated speck-like protein endowed with a CARD domain; BMAA = β-methylamino-L-alanine; CASP1 = caspase-1; CD14 = cluster of differentiation 14; GSDMD-N = gasdermin D N-terminal fragments; NF-κB = nuclear factor κB; NOX2 = NADPH oxidase 2; P2X7R = purinergic receptor; ROS = reactive oxygen species; SOD1 = superoxide dismutase 1; TDP-43 = TAR DNA-binding protein 43; TLR− =Toll-like receptor−; TNF-α = tumor necrosis factor-α; WT = wild-type. The arrows show the sequences of molecular events induced by stressors and factors.
Biomedicines 11 00999 g004
Table 1. Main conditions and factors activating the brain NLRP3 inflammasome.
Table 1. Main conditions and factors activating the brain NLRP3 inflammasome.
Condition/FactorMechanismsReferences
Vascular ailments
Stroke
Intracerebral hemorrhage
Hemorrhagic stroke
Mitochondrial disfunction after hypoxic
ischemia/reperfusion (HI/R)
Chronic hypoxia
[32,33,34,35,36,37,38,39,40]
Seizures
Mesial lobe temporal epilepsy
Soman or A255 (nerve agent) exposure
Acetyl- and butyryl-cholinesterase inhibition[41,42]
Metal accumulation
Manganese (Mn),
Lead (Pb)
Copper (Cu)
Cadmium (Cd)
Aluminium/alum
Metal-induced neurotoxicity
§ ROS & NF-κB-p65 pathway
CaSR and GCP6RA signaling
[43,44,45,46,47,48,49,50,51]
See also
Box 1
Mechanical stresses and strains
Skull trauma
Optic nerve trauma
Elevated intracranial pressure
Glaucoma

Osteopontin
NIMA-related kinase 7 (or NEK7)
P2X7 receptor activation
HMGB1/caspase-8 pathway
[52,53,54,55,56,57,58,59]
See also
Box 2
Neurodegenerative diseases
Alzheimer’s disease (AD)
Tauopathies
Parkinson’s disease (PD)
Amyotrophic lateral sclerosis (ALS)
Huntington’s disease (HD)
Prion disease (PrPSc)


Aβs, autophagy block, NEK7
p-Taues paired helical filaments
ER stress, ↑ ROS
α-Synuclein aggregates
Mutated SOD1, TDP-43
Expanded CAG repeats in HTT/OT15 gene
Prion protein seeding
[60,61,62,63,64,65,66,67,68,69,70,71]
Environmental pollution
PM2.5
Increased ROS production by microglia[72,73]
Infectious diseases
Sepsis (bacteria, fungi)
West Nile Virus (WNV)
HIV-1
Herpes Virus 1
Japanese Encephalitis Virus (JEV)
Zika Virus (ZIKV)
SARS-CoV-2
Encephalomyocarditis Virus (EMCV)
Tuberculosis


Bacterial and fungal toxins
Intensified IL-1β signaling
Tat and gp120 proteins
Gasdermin D-dependent
pyroptosis
ROS-dependent activation of Src/Ras/Raf/ERK/NF-κB
signaling axis
NS5 protein and ↑ ROS
S1 spike glycoprotein, viroporin ORF3a/8
viroporin ORF2b
Early secreted antigenic target protein of 6 kDa (ESAT-6)
[74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90]
Metabolic disorders
Atherosclerosis
Gout
Obesity/high-fat diet
Nonalcoholic hepatic steatosis
Type-2 diabetes mellitus (T2DM)
Hypercholesterolemia
Urates→NEK7
Glucocorticoids and fatty acids surpluses→TNFR and Toll-like receptors
ROS, NO, hydroperoxides, scavenger receptors, mTOR
[91,92,93,94]
Iatrogenic factors
Postoperative cognitive dysfunction
Cyclophosphamide cystitis
GdCl3, cinacalcet
Glucocorticoids (elevated levels)
Drugs, infection, electrolyte imbalance
TNF-α
Calcimimetic•CaSR/ERK1/2/CaMKII
NLRP1 and NLRP3 inflammasomes
[95,96,97,98,99,100,101]
See also Box 1
Psychotropic drugs
Cocaine
Methamphetamine
Scopolamine
Ethanol
Morphine
Fentanyl
σ-1 receptor
TLR-4
↑ Dhx58, S100a, Lrm4 genes
TLR-4
μ-3 and κ opioid receptors
μ-opioid receptor
[102,103,104,105,106,107,108]
Cellular stress and injury
ATP
Pore-inducing agents
Phagocytosed protein polymers
ROS
Cardiolipin
Raised IL-1β levels
Reduced cyclic AMP (cAMP) levels
Zn2+ deficiency
K+ efflux
Ca2+ and Cl influx
Purinergic receptor signaling
DDX3X protein/NLRP3 complexes
Heat shock protein 60 (HSP60) and
TLR-4-p38 MAPKs axis
Oxidized mtDNA and proteins
Lysosome-released cathepsin B
Mitochondria-released hexokinase, ROS
NLRP3 activation
Ionic imbalances
[25,109,110,111,112,113,114,115,116,117,118,119]
Aging
Inflammaging
↑ Membrane attack complexes (MAC)
Reduced mitochondrial fission and fusion
Declined mitophagy
Mitochondrial damage
Selective autophagy-mediated mitochondrial
homeostasis (in microglia)
[33,112,120,121,122]
§ ↑ = increased.
Table 2. RNAs modulating brain NLRP3 inflammasome’s function.
Table 2. RNAs modulating brain NLRP3 inflammasome’s function.
(A) Activation.
RNAsModelMechanismsReferences
LncRNA-Cox2Murine microglia↑ § Transcription of NLRP3 and ASC
TLR-mediated signaling
pathways
Autophagy block
Microglia activation
[180,181,188]
LncRNA-Meg3Murine microgliamiR-7a-5 downregulation[189]
miR-141Brain tissue of diabetic miceNF-ĸB-mediated
NLRP3 expression
[190]
Exo-miR-124
Exo-miR-146a
Exo-miR-155
LPS-primed N9 microglia cells↑ TLR4/TLR2/NF-ĸB axis[191]
miR-193Murine brain cortex
Murine microglia
↑ Expression of NLRP3, ASC, cleaved caspase-1 and mature
IL-1β
[192]
miR-590-3In silico AD patients’ dataPromoted neurons’ death via AMPK signaling [193]
P3Alu-RNAsPrimary human retinal pigment cells ERK1/2 and NLRP3 activation, neurons’ death[184]
(B) Inhibition.
RNAsModelMechanismsReferences
circRNA_003564Spinal cord injury (rat model)↓ § NLRP3, caspase-1,
mature IL-1β, Il-18, GsdmD
↓ Pyroptosis
[187]
LncRNA-Meg3Rat hippocampal neuronal model of temporal epilepsyPI3K/AKT/mTOR
pathway activation
[194]
miR-7Murine neural stem cellsNLRP3/caspase-1
suppressor
[195,196]
Exo-miR-21APP/PS1 2xTg AD-model mouseImproved memory [197]
miR-22, Exo-miR-22APP/PS1 2xTg AD-model mouse
PC12 cells
Downregulated NLRP3[198,199]
Exo-miR-23bRat model of intracerebral hemorrhageAntioxidant effects via PTEN/NRF2 inhibition[200]
miR-29c-3p
Exo-miR-29c-3p
PC12 cells
AD-model rat
Suppression of BACE1,
p-Tau, and pyroptosis via Wnt/β-catenin pathway
[201,202]
miR-152Microglial BV2 cell
Hippocampal neuronal HT22 cell line
Rat model of intracerebral hemorrhage
TXNIP-mediated block of NLRP3 activation[203]
Exo-miR-188-3p PD-model mouse
MN9D dopaminergic neuronal cells
Suppression of
NLR3/pyroptosis
[204]
miR-194-5pRat model of intra-
cerebral hemorrhage
Blocked NLRP3/TRAF6
interaction
[205]
miR-223-3pSerum samples from PD, AD, and MCI patients, and healthy controls Negative NLRP3
regulation
[206]
miR-374a-5pRat model of hypoxic-ischemia encephalopathySuppressor of SMAD6/NLRP3
in microglia
[207]
§ ↑ = increased; ↓ = decreased.
Table 3. Inhibitors of brain NLRP3 inflammasome.
Table 3. Inhibitors of brain NLRP3 inflammasome.
Compound
[References]
IUPAC NameMain Molecular
Activity
Main Biological
Activity
Experimental Model
17β-Estradiol (E2)
[223,224,225]
See also Box 1
(8R,9S,13S,14S,17S)-13-methyl-6,7,8,9,11,12,14,15,16,17-decahydrocyclopenta[a]phenanthrene-3,17-diolLigand for estrogen receptor-α (ER-α) and -β (ER-β), and for G-protein coupled receptor 1 (GPER1)§ NLRP3, ASC, cleaved caspase-1, IL-1β
↓§ M1 microglia
↑ M2 microglia
Male SOD1(G93A) ALS-model mice
Global brain ischemia-model rodents
A43879
[226]
See also Box 2
3-[[5-(2,3-dichlorophenyl)-tetrazol-1-yl]methyl]pyridine hydrochlorideP2X7 purinergic receptor antagonist↓ P2X7 receptor signaling
↓ NLRP3
Spinal cord injury-model animal
Adiponectin
[227]
ProteinLigand for Adipo-R1 and Adipo-R2 receptors↓ NLRP3, IL-1β, IL-18
↑ Autophagy via AMPK
pathway
Intracerebral
hemorrhage-model rat
Amifostine
[228]
2-(3-aminopropylamino)ethyl-sulfanylphosphonic acidProtects against the DNA-damaging effects of ionizing radiations and chemotherapy drug-induced ROS↓ ROS, pyroptosisExperimental autoimmune encephalomyelitis (EAE)-model rat
α1-Antitrypsin (A1AT)
[128]
ProteinProtease inhibitor↓ Aβ1–42-driven NLRP3
activation
Mouse primary
cortical astrocytes
Anfibatide
[229,230]
Dimeric proteinAntagonist of the glycoprotein Ib IX-V (GPIb)
complex
↓ NLRP3/NF-κB axis, cleaved caspase-1 and -3, and Bax
↑ Bcl2
Cerebral HI/R injury-model rat
Atorvastatin
[231]
(3R,5R)-7-[2-(4-fluorophenyl)-3-phenyl-4-(phenylcarbamoyl)-5-propan-2-ylpyrrol-1-yl]-3,5-dihydroxyheptanoic acidInhibitor of
hydroxymethylglutaryl-coenzyme A (HMG-CoA) reductase
↓ NLRP3/NF-κB signaling axisSurgery-induced BBB disruption in aged mice
Bay117082
[232]
(E)-3-(4-methylphenyl)-sulfonylprop-2-enenitrileCalcium channel blocker↓ ATPase activity of NLRP3Spinal cord injury-model animal
BPBA
[233]
(2-[2-(benzo[d]thiazol-2-yl) phenyl-amino] benzoic acid)Inhibitor of self- and Cu2+- or Zn2+-induced Aβs aggregation↓ Aβs aggregation and
neurotoxicity
↓ NLRP3 and IL-1β
Aβ-induced paralysis in transgenic
Caenorabditis elegans
Caffeine
[234]
1,3,7-trimethypurine-2,6-dioneAntagonist of all adenosine receptor subtypes (A1, A2a, A2b, A3)
in the CNS PDE inhibitor
↓ Rapamycin (mTOR) axis and Bax
↑ Autophagy
EAE-model C57BL/6 mice
Mouse microglia
BV2 microglial cells
Calcitriol
[235]
(1R,3S,5Z)-5-[(2E)-2-[(1R,3aS,7aR)-1-[(2R)-6-hydroxy-6-methylheptan-2-yl]-7a-methyl-2,3,3a,5,6,7-hexahydro-1H-inden-4-ylidene]ethylidene]-4-methylidenecyclohexane-1,3-diolLigand for vitamin D
receptors
↓ ROS, NLRP3, caspase-1, IL-1β, CX3CR1, CCL17, Tbx21
↓ Spinal cord
demyelination
EAE-model C57BL/6 mice
Choline
[236]
2-hydroxyethyl-(trimethyl)azaniumMethyl donor
Ligand for choline transporters, CTL1 included
↓ NLRP3, Aβs deposition, and
microgliosis
APP/PS1 AD-model mice
Dapansutrile
(i.e., OLT1177)
[237]
3-methylsulfonyl
propanenitrile
Direct NLRP3 ATPase inhibitor↓ Microglia activation and Aβs plaque numbers in the cerebral
cortex
↓ IL-1β and IL-6
↑ Dendritic spine density
Successful Phase I clinical trial
APP/PS1 AD-model mice
Dexmedetomidine (Dexm)
[96,238]
5-[(1S)-1-(2,3-dimethylphenyl)ethyl]-1H-imidazoleSpecific and selective α-2
adrenoceptor agonist
↓ NF-κB and proinflammatory cytokines via miR-340 upregulation
↑ Autophagy
LPS-stimulated BV2 microglia cells
Dihydromyricetin
[239]
(2R,3R)-3,5,7-trihydroxy-2-(3,4,5-trihydroxyphenyl)-2,3-dihydrochromen-4-oneAntioxidant, anti-binge hangover, and anti-cancer activity↓ NLRP3
↑ Aβ clearance
↑ Expression of neprilysin
↑ M2 microglial phenotype
APP/PS1 AD-model mice
A-68930
[240]
1-(aminomethyl)-3-phenyl-3,4-dihydro-1H-isochromene-5,6-diol;hydrochloridePotent and selective
Dopamine D1-like receptor agonist
↓ NLRP3 activationLPS-induced systemic
inflammation mouse model
Bromocriptine(6aR,9R)-5-bromo-N-[(1S,2S,4R,7S)-2-hydroxy-7-(2-methylpropyl)-5,8-dioxo-4-propan-2-yl-3-oxa-6,9-diazatricyclo[7.3.0.02,6]dodecan-4-yl]-7-methyl-6,6a,8,9-tetrahydro-4H-indolo[4,3-fg]quinoline-9-carboxamideDopamine D2 receptor
agonist
↑ NLRP3 ubiquitination via cAMPNeurotoxin MPTP-treated mice
Dopamine
[226]
4-(2-aminoethyl)benzene-
1,2-diol
Agonist for the five
Dopamine receptor
subtypes (D1, D2, D3, D4, D5)
↓ IL-1β and IL-18 secretionSpinal cord injury-model rat
LY171555(4aR,8aR)-5-propyl-1,4,4a,6,7,8,8a,9-octahydropyrazolo[3,4-g]quinoline;hydrochlorideSpecific dopamine D2
receptor agonist
Quinerolane(5aR,9aR)-6-propyl-5a,7,8,9,9a,10-hexahydro-5H-pyrido[2,3-g]quinazolin-2-amineDopamine D2 and D3 receptors
agonist
EC144
[241]
5-[2-amino-4-chloro-7-[(4-methoxy-3,5-dimethylpyridin-2-yl)methyl]pyrrolo[2,3-d]pyrimidin-5-yl]-2-methylpent-4-yn-2-olSelective inhibitor of heat shock protein 90 (HSP90)↓ IL-1β and IL-18Peritonitis-model
animal
Echinacoside
[242]
[(2R,3R,4R,5R,6R)-6-[2-(3,4-dihydroxyphenyl)ethoxy]-5-hydroxy-2-[[(2R,3R,4S,5S,6R)-3,4,5-trihydroxy-6-(hydroxymethyl)oxan-2-yl]oxymethyl]-4-[(2S,3R,4R,5R,6S)-3,4,5-trihydroxy-6-methyloxan-2-yl]oxyoxan-3-yl] (E)-3-(3,4-dihydroxyphenyl)prop-2-enoateNeuroprotective effects via
undefined upstream
mechanisms
↓ NLRP3, NF-κB-p65, and ROSSpinal cord injury-model animal
LPS-treated BV2
microglial cells
Ellagic acid
[243]
6,7,13,14-tetrahydroxy-2,9-dioxatetracyclo[6.6.2.04,16.011,15]hexadeca-1(15),4,6,8(16),11,13-hexaene-3,10-dioneATP-competitive inhibitor of
constitutively active CK2 Ser/Thr protein kinase
↓ caspase-1, IL-6, IL-10, IL-17A,
TNF-α, GFAP, and Iba1
EAE-model mouse
Fimasartan
[244]
2-[2-butyl-4-methyl-6-oxo-1-[[4-[2-(2H-tetrazol-5-yl)
phenyl]phenyl]methyl]pyrimidin-5-yl]-N,N-dimethylethanethioamide
Angiotensin II receptor
antagonist
↓ NLRP3/ASC/caspase-1 and NF-κB pathwaysIntracerebral hemorrhage-model rat
Hemolysate-treated BV2 microglia
Fluoxetine
[245]
N-methyl-3-phenyl-3-[4-(trifluoromethyl)phenoxy]propan-1-amineSerotonin reuptake
inhibitor
↓ NF-κB, TLR-4, NLRP3, caspase-1, TNF-α, IL-1β
↓ AChE activity, Aβ, Tau protein, MDA
Depression- and AD-model animals
Ghrelin
[246]
(4S)-4-[[(2S)-1-[(2S)-2-[[(2S)-2-[[(2S)-2-[[(2S)-2-[[(2S)-2-[(2-aminoacetyl)amino]-3-hydroxypropanoyl]amino]-3-hydroxypropanoyl]amino]-3-phenylpropanoyl]amino]-4-methylpentanoyl]amino]-1-oxopropan-2-………………………………….
yl]amino]-5-oxopentanoic acid
Ligand for GHS-R1a
receptor
↓ NF-κB/NLRP3 axis, IL-6, COX2, TNF-α, NOS-2, and pyroptosisEAE-model animal
Glibenclamide
[247,248]
5-chloro-N-[2-[4-(cyclohexylcarbamoylsulfamoyl)phenyl]ethyl]-2-methoxybenzamideClassic KATP channel
blocker
ATP-sensitive K+ channel inhibitor
↓ NLRP3
↓ Release of HSP70

↓ NLRP3, GsdmD-cleavage,
↓ Oxidative stress, demyelination, axon degeneration
Morphine-induced
neuroinflammation
animal and cellular models

Hexanendione-induced neurotoxicity-model animal
HU-308
[249]
[(1R,4R,5R)-4-[2,6-dimethoxy-4-(2-methyloctan-2-yl)phenyl]-6,6-dimethyl-2-bicyclo[3.1.1]hept-2-enyl]methanolActivator of cannabinoid receptor 2↑ AutophagyBV2 microglia cells
EAE-model animals
Indomethacin
[250]
2-[1-(4-chlorobenzoyl)-5-methoxy-2-methylindol-3-yl]acetic acidProstaglandin G/H
synthase 2 or cyclo-oxygenase (COX) enzyme inhibitor
↓ NLRC4 and NLRP3 genes
↓ IL-1β, caspase-1, and
p-Taues
Streptozotocin (STZ)-induced AD-like model
Inzomelid
[251]
1-(1,2,3,5,6,7-hexahydro-s-indacen-4-yl)-3-(1-propan-2-ylpyrazol-3-yl)sulfonylureaNonspecific and reversible inhibitor of the cyclo-oxygenase (COX)
enzyme or prostaglandin G/H
synthase
↓ NLRP3ClinicalTrial.gov NCT04015076
JC124
[252]
5-chloro-2-methoxy-N-[2-[4-(methylsulfamoyl)phenyl]ethyl]benzamide)Specific inhibitor of expression of NLRP3 and its adaptor protein ASC↓ NLRP3, ASC, IL-1β, TNFα, NOS-2, caspase-1, and pyroptosisTraumatic brain injury in male rats
Ketamine
[253]
2-(2-chlorophenyl)-2-(methylamino)cyclohexan-1-oneNMDA receptors
antagonist
↓ NF-κB, NLRP3, ASC, caspase-1,
IL-1β
↑ Autophagy
Depressive-like-model rat
KPT-8602
[254]
(E)-3-[3-[3,5-bis(trifluoromethyl)phenyl]-1,2,4-triazol-1-yl]-2-pyrimidin-5-ylprop-2-enamideExportin 1 (XPO1) nuclear transport inhibitor↓ Exportin 1
↓ NLRP3/NF-κB signaling axis
LPS-treated
macrophages
LPS-induced inflammation mouse model
MPTP mouse model of PD
Licochalcone B
[255]
(E)-3-(3,4-dihydroxy-2-methoxyphenyl)-1-(4-hydroxyphenyl)prop-2-en-1-oneSpecific inhibitor of NEK7-NLRP3 interaction↓ Canonical and
non-canonical NLRP3
inflammasome activation
Murine macrophages
Mouse models of LPS-induced septic shock, peritonitis, and non-alcoholic steatohepatitis
Manoalide
[256,257,258,259]
(2R)-2-hydroxy-3-[(2R,6R)-6-hydroxy-5-[(E)-4-methyl-6-(2,6,6-trimethylcyclohexen-1-yl)hex-3-enyl]-3,6-dihydro-2H-pyran-2-yl]-2H-furan-5-oneInhibitor of NEK7-NLRP3
activating interaction
↓ Canonical and
non-canonical NLRP3
inflammasome activation
EAE-model animal
MCC950
(i.e., CRID3)
[222,260]
1,2,3,5,6,7-hexahydro-s-indacen-4-ylcarbamoyl-[4-(2-hydroxypropan-2-yl)furan-2-yl]sulfonylazanideSelectively and specifically binds NLRP3 NATCH domain hindering Walker B motif function thereby
inhibiting NLRP3 conformational modifications and oligomerization
↓ NLRP3
↑ Aβ-phagocytic capability of
microglia
↓ IL-1β, IL-18, TNF-α, NLRP3, ASC, cleaved caspase-1, Iba1-, and
GFAP-positive cells
↑ BDNF and PSD95 expression
APP/PS1 transgenic AD-model mouse
LPS + ATP-induced
microglia
Perioperative neurocognitive disorders-model mice
Mefenamic,
Tolfenamic,

Flufenamic,

Meclofenamic acids
[261]
2-(2,3-dimethylanilino)benzoic acid
2-(3-chloro-2-methylanilino)benzoic acid
2-[3-(trifluoromethyl)anilino]benzoic acid
2-(2,6-dichloro-3-methylanilino)benzoic acid
Cyclooxygenase (COX)
inhibitors
Cl- channel inhibitors
↓ NLRP3 and IL-1β
processing and release
LPS-primed primary bone
marrow-derived macrophages
Melatonin
[262,263,264,265,266]
N-[2-(5-methoxy-1H-indol-3-yl)ethyl]acetamideNatural hormone of the pineal gland acting through its receptors↑ TFEB nuclear
translocation
↑ mitophagy
↓ NLRP3, IL-18, IL-6, and IL-1β
↓ ROS
↑ Sirtuin 1
↑ α7-nAChR-mediated
“autophagic flux”
25–35-treated SH-SY5Y cells
APP/PS1 AD-model mice
Chronic Gulf War
syndrome
Metformin (MET)
[267]
3-(diaminomethylidene)-1,1-dimethylguanidineAMP-activated protein
kinase (AMPK) agonist
↓ NF-κB signaling pathway
↑ Sirtuin 1
↓ NLRP3-mediated ECs pyroptosis
LPS-stimulated lung
tissues and pulmonary endothelial cells
Milrinone
[268]
6-methyl-2-oxo-5-pyridin-4-yl-1H-pyridine-3-carbonitrileInhibitor of
phosphodiesterase III
↑ cAMP
↓ TLR4/MyD88/NF-κB axis
↓ IL-1β, IL-6, TNF-α
↓ Aβ, p-Tau, ROS
LPS/Aβ-treated BV2 microglial cells
APP/PS1 AD-model mouse
Minocycline
[269,270]
(4S,4aS,5aR,12aR)-4,7-bis(dimethylamino)-1,10,11,12a-tetrahydroxy-3,12-dioxo-4a,5,5a,6-tetrahydro-4H-tetracene-2-carboxamideCaspase-1 negative
modulator
↓ TLR-2, MyD88,
NLRP3/NF-κB axis, IL-1β
AD-like dementia-model mouse
Mitoquinone
(MitoQ)
[271]
10-(4,5-dimethoxy-2-methyl-3,6-dioxocyclohexa-1,4-dien-1-yl)decyl-triphenylphosphaniumSelectively accumulates
inside mitochondria with
anti-oxidant action
↓ Mitochondrial ROS, NLRP3
activation, IL-1β, and IL-18
↑ M2 phenotype microglia
Intracerebral
hemorrhage-model mouse
FeCl2-treated microglia
N-acetylcysteine
[272]
(2R)-2-acetamido-3-sulfanylpropanoic acidStimulator of glutathione synthetase ↓ ROS
↑ NRF2-induced NAD(P)H quinone dehydrogenase 1 (NQO1)
Ischemic stroke-model rat
Nafamostat
mesylate
[273]
(6-carbamimidoyl naphthalen-2-yl) 4-(diaminomethyl-ideneamino)benzoateSynthetic inhibitor of
serine proteases
with a wide spectrum of activity
↓ NLRP3/NF-κB signaling
↓ TNF-α, IL-1β, NOS-2, COX-2, IL-18
Stroke-model animal
NT-0796
[274]
unknownOrally available brain-
penetrant NLRP3 inhibitor
↓ NLRP3ANZCTR.org.au ACTRN126210010828-97
Phenyl vinyl
sulfone
[275]
ethenylsulfonylbenzeneCysteine protease inhibitor↓ NLRP3-mediated IL-1β releaseLPS+ATP-treated J774A.1 cells
LPS intraperitoneally injected C57BL/6 mouse
Phoenixin-14
[276]
See also Box 1
proteinLigand for the multiple function G protein-coupled receptor GPR173↓ HMGB1-mediated NLRP3
activation
↓ IL-1β and IL-18
LPS-treated mouse
primary astrocytes
Pramipexole
[277]
6S)-6-N-propyl-4,5,6,7-tetrahydro-1,3-benzothiazole-2,6-diamineDopamine-D3 receptors agonist↑ Autophagy
↓ NLRP3, ASC, cleaved caspase-1 IL-1β, IL-18
LPS+ATP-stimulated primary mouse astrocytes
PD-model mouse
Prednisone (PDN)
[278]
(8S,9S,10R,13S,14S,17R)-17-hydroxy-17-(2-hydroxyacetyl)-10,13-dimethyl-6,7,8,9,12,14,15,16-octahydrocyclopenta[a]phenanthrene-3,11-dioneGlucocorticoid receptor agonist↓ NLRP3 activation
↓ TNF-α, CCL8, CXCL10, CXCL16
↓ astrocytes and microglia activation
Cuprizone (CPZ)-
induced demyelination-model mouse
Resolvin D1
[279]
See also Box 1
(4Z,7S,8R,9E,11E,13Z,15E,17S,19Z)-7,8,17-trihydroxydocosa-4,9,11,13,15,19-hexaenoic acidLigand for N-formyl
peptide receptor-2 and
GPR-32
↑ A20 expression
↓ NLRP3/NF-κB axis
Subarachnoid
hemorrhage-model rat
Sildenafil
[280]
5-[2-ethoxy-5-(4-methylpiperazin-1-yl)sulfonylphenyl]-1-methyl-3-propyl-6H-pyrazolo[4,3-d]pyrimidin-7-one3′,5′-cyclic GMP (cGMP)-specific phosphodiesterase
inhibitor
↓ NLRP3
↓ Hippocampal Aβ1–40 and Aβ1–42
↑ Brain cGMP levels
APP/PS1 AD-model mouse
TAK-242
(CLI-095)
[103,281]
(R)-Ethyl 6-(N-(2-chloro-4-fluorophenyl)sulfamoyl)cyclohex-1-enecarboxylateTLR-4 signal transduction inhibitors↓ TLR-4-NF-κB-caspase-11 axis
↓ NLRP3, IL-1β, and IL-18
Methamphetamine-treated mouse and
primary astrocytes
1–42-treated BV2 microglia and HT-22 neurons
1,2,4-TTB
[282]
1,2,4-TrimethoxybenzeneInhibitor of NLRP3
oligomer formation
↓ Nigericin- or ATP-mediated NLRP3 activationMurine bone marrow-derived macrophages (BMDMs)
Primary mouse
microglia
EAE-model mice
Urolithin A
[283]
3,8-dihydroxybenzo[c]chromen-6-oneGut microflora processed
derivative of ellagic acid
↓ NLRP3 activation via
mitophagy promotion in
microglia
LPS- or MPTP-treated BV2
microglial cells
MPTP PD-model mouse
VX-765
[284]
(2S)-1-[(2S)-2-[(4-amino-3-chlorobenzoyl)amino]-3,3-dimethylbutanoyl]-N-[(2R,3S)-2-ethoxy-5-oxooxolan-3-yl]pyrrolidine-2-carboxamideCompetitive inhibitor of ICE/caspase-1 (active
metabolite: VRT-043198)
↓ NLRP3/caspase-1/GsdmD pathwayAPP/PS1 AD-model mice
BV2 microglial cells
§ ↑ = increased; ↓=decreased.
Table 4. Brain NLRP3 inflammasome downregulation by officinal plants agents/extracts.
Table 4. Brain NLRP3 inflammasome downregulation by officinal plants agents/extracts.
(A) Agents.
Natural Compounds and SourcesChemical ClassBiological ActivitiesExperimental ModelsReferences
Andrographolide
from the roots and leaves of the plant Creat or Green chireta (Andrographis paniculata Wall. ex Nees)
labdane diterpenoid↓§ P2X7 receptor signaling
↓ HMGB1-induced TLR-4-NFκB
signaling
LPS-activated mixed glial cells
LPS-treated mouse
[285]
see also Box 2
Artesunate/Artemisinin
from Artemisiae Iwayomogii Herba
sesquiterpene lactone↓ Inflammatory response and neuron death
§ Expression of BDNF, GDNF, and NT-3 neurotrophins
Traumatic brain injury-model mouse
LPS-stimulated BV-2 microglial cells
LPS-treated mouse
[286,287]
Astragaloside IV
from Astragalus membranaceus (i.e., Huangqi)
pentacyclic triterpenoidAntioxidant activityTransient cerebral ischemia/reperfusion (I/R)-model mice[288]
Baicalin
from the root of Scutellaria baicalensis Georgi
flavonoid↓ TLR-4/NF-κB/NLRP3 axisAPP/PS1 AD-model mice
LPS/Aβ-stimulated BV2 microglial cells
[289]
Benzyl isothiocyanate
from cruciferous vegetables
benzene↓ NLRP3 activation via mitochondria-
generated ROS inhibition
↓ NF-κB signaling
LPS-induced BV2 microglial cells[290]
Bixin
from the seeds of the Achiote tree (i.e., Bixa orellana)
apocarotenoidSuppression of thioredoxin-interacting protein (TXNIP)-NLRP3 activityEAE-model mouse [291]
Carnosic acid (CA)
Carnosol (CS)
from Rosmarinus officinalis
abietane-type tricyclic diterpenes↑ KEAP1 (Kelch-like ECH-associated protein 1)/NRF2 (erythroid 2–related factor 2) transcriptional pathway
activation
↓ HSP 90 inhibition
APP/PS1 AD-model mice
Primary mouse bone marrow-derived
macrophages
[292,293]
Cucurbitacin B
from Cucurbitaceae
tetracyclic triterpene↓ NLRP3, caspase-1 self-activation,
and IL-1β release
Ischemia/reperfusion injury-model rat [294]
Dehydroisohispanolone
diterpene (DT1)
from Ballota hispanica (Labiatae)
labdane (bicyclic diterpene)↓ NF-κB and NLRP3 signalingNigericin-activated murine bone
marrow-derived macrophages
[295]
Demethylene-tetrahydroberberine (DMTHB)
from Berberis vulgaris, Berberis
aristata
alkaloid↓ NLRP3 inflammasome’s activation
↓ IL-6 signaling
AD-model mice[296]
Esculentoside A
from the roots of Indian pokeweed (i.e., Phytolacca esculenta Van Houtte)
triterpene saponin↓ NF-κB, MAPKs and NLRP3
pathways
LPS-activated murine primary
microglia cells and BV2 microglia cells
[297]
Gastrodin
from rhizome of Gastrodia elata Blume
phenolic glycoside↓ TLR4-NF-κB-NLRP3 axis and
microglia-mediated
neuroinflammation
LPS-treated rats[298]
Ginkgolide B
(BN-52021)
from Ginkgo biloba and
Machilus wangchiana
diterpenoid esters ↓ NLRP3 and microglia-mediated
neuroinflammation
↑ NLRP3 autophagic degradation
1–42-induced BV2 cells
LPS-primed BV2 cells
senescence-accelerated male mouse prone 8 (SAMP8)
[299,300]
Ginsenosides
(Rb1, Rg1, Rg3, Rg5, Rh1, Compound K, Chikusetsusaponin IVa, Gintonin,
and 20(S)-Protopanaxatriol)
from Panax ginseng C.A. Meyer; Panax quinquefolius L. (i.e., American Ginseng); and Panax japonicus T. Nees
saponins↓ NLRP3, NLRP1, AIM-2, and caspase-1 self-activation
↓ brain load of Aβs
↑ soluble (s)APP-α
AD in rodent models
Depression-like behavior in rat model
Post-traumatic stress disorder-like
behavior in rodent model
Stroke model
High fat diet-model mouse
[301,302,303,304,305]
Isoformononetin
from Cicer arietinum L. (chickpea)
methoxyisoflavone↓ NLRP3, NLRP2, ASC, NFκB-p65, IL-1β, caspase-1 proteins, and ROSStreptozotocin-treated rat[306]
Isoliquiritigenin
from the Chinese herbal
medicine Glycyrrhiza (Guo Lao)
isoflavone↓ NLRP3
↑ NRF2-induced antioxidant activity
Hippocampal organotypic slice cultures
after oxygen/glucose deprivation (OGD)
[307]
Isosibiricin
from orange jasmine (i.e., Murraya exotica or paniculata)
coumarinNLRP3-inhibition mediated by
Dopamine D1/2 receptors
LPS-primed mouse BV-2 microglial cells[308]
Kaempferol
from several herbs in TCM
polyphenol flavonoid↑ NLRP3 autophagic degradationPD-model mouse
LPS-primed BV-2 microglial cells
[309,310,311]
β-Lapachone
from the Lapacho tree or
Jacaranda (i.e., Tabebuia Avellaneda Lorentz)
benzochromenoneAntioxidant activityMultiple sclerosis and
AD-model animals
[312]
Lychee seed polyphenols (LSPs)
from the Litchi chinensis tree
polyphenols ↑ Autophagy via the AMPK/mTOR/ULK1 axis
↑ Tight junctions’ expression
↑ LRP1 (i.e., low-density lipoprotein
receptor-related protein 1),
Beclin 1, and LC-3II proteins
Aβ-induced BV2 microglia cells
APP/PS1 AD-model mouse
[313,314]
Mangiferin
from the rhizome of Anemarrhena asphodeloides Bunge
C-glucoside
xanthone
↓ NF-κB and NLRP3 signaling
↓ Microglial M1 polarization
LPS-induced BV2 cells[315]
Myricitrin
from the root bark of the tallow shrub (i.e., Myrica cerifera L.)
polyphenol hydroxy flavonoid↓ NLRP3/Bax/Bcl2 axis
NF-κB inactivation
Antioxidant activity
Rat model of sepsis-linked
encephalopathy
Brain HI-model rat
[316,317]
Neferine
from the green seed embryos of the lotus plant (
i.e., Nelumbo nucifera Gaertn)
bisbenzylisoquinoline alkaloid↓ NLRP3-mediated neuronal
pyroptosis
Neonatal HI brain
damage model rat
PC12 cells
[35]
Nobiletin
from Citrus L. fruits
polymethoxylated flavonoid↓ NLRP3
↑ Autophagy via AMPK/mTOR/ULK1 axis
LPS-treated rat brain and BV2 cells[318]
Oleocanthal
from extra-virgin olive oil
phenylethanoid↓ NLRP3
↑ Autophagy via AMPK/mTOR/ULK1 axis
AD-model TgSwDI Mouse [319]
Oridonin
from Isodon Rubescens
(Hemsl.) H. Hara
(1S,2S,5S,8R,9S,10S,11R,15S,18R)-9,10,15,18-tetrahydroxy-12,12-dimethyl-6-methylidene-17-oxapentacyclo[7.6.2.15,8.01,11.02,8]octadecan-7-oneBinds NLRP3’s NACHT domain blocking NEK-7-NLRP3
activating interaction
↓ NF-κB pathway, Aβ1–42-elicited
neuroinflammation, and pyroptosis
1–42-induced AD mice[320]
Osthole
from the roots of various medicinal plants, including Cnidium monnieri L. and Angelica pubescens
(Japan’s Shishiudo).
7-methoxy-8-(3-methylpent-2-enyl) coumarin↓ NLRP3
↓ brain load of Aβs
Rat model of chronic cerebral ischemic
hypoperfusion
[321]
Purpurin
from Rubia tinctorum L.
Rhein
from Rheum rhabarbarum
anthraquinones↓ NLRP3, caspase-1 self-activation, and IL-1β releaseAD-model animals
Perirhinal cortex high-fat-diet-induced
animal model
[322]
Quercetin
(plant pigment)
flavonoidAntioxidant activity
↓ NLRP3-pyroptosis-mediated
IL-1β release
↑ Sirtuin
LPS-induced primary microglial cells and BV2 cells
LPS-induced PD model mouse
Depression-model mouse
SAMP8 mice
[323,324]
Sinomenine
from the roots of the climbing plant Sinomenium acutum (Thumb.)
alkaloidAntioxidant and anti-inflammatory
activity
EAE-model mouse[325]
Thonningianin A
from Penthorum chinense
ellagitannin
polyphenol
↑ NLRP3 autophagic degradation via AMPK/ULK1 and Raf/MEK/ERK axisIn vitro and in vivo AD models,
including, C. elegans, APP/PS1 mice,
BV-2 cells, and PC-12 cells
[119]
Withaferin
from Indian ginseng
(i.e., Withania somnifera)
steroidal lactone↓ Gene expression of NF-κB and associated neuroinflammatory molecules SH-SY5Y cells transfected with APP plasmid (SH-APP)[326]
(B) Herbal Extracts.
Herbal/Fruit ExtractSourceBiological ActivityExperimental ModelReferences
Açaí extractBerries of the Euterpe oleracea Mart. palm treeAntioxidant activityLPS- or nigericin-activated
microglia (EOC 13.31) cells
[327]
Crysanthemum indicum
extract (CIE)
TCM (main components: chlorogenic acid, luteoloside, and 3,5-dicaffeoylquinic acid)Antioxidant activity
↑ TrkB/Akt/CREB/BDNF and
Akt/Nrf-2/ARE axes
H2O2-induced oxidative toxicity in
hippocampal HT22 neuronal cell line
[328,329]
Glycyrrhiza
(Guo Lao)
TCM (main components:
licochalcone, isochalcone A, echinatin, isoliquiritigenin, and glycyrrhizin)
↓ NLRP3, TNF-α, IL-1β, and IL-18
↑ AMPK/NRF2/antioxidant response
element (ARE) signaling
LPS-induced chondrocyte pyroptosis
LPS-induced macrophage cells
Ischemic brain damage-model animal
[307,330]
KutkiAyurvedic medicine
from rhizomes and roots of Picrorhiza kurroa
↓ NLRP3 and BACE-1 expression5xFAD-model mice[331]
Pien-Tze-HuangTCM, including Radix et Rhizoma Notoginseng, Moschus, Calculus Bovis, and Snake Gall ↓ NLRP3
↑ Autophagy via AMPK/mTOR/ULK1 axis
LPS-induced BV2 microglial cells
cerebral ischemia/reperfusion impaired rats
[332]
TojaprideTCM (main components: Cyperus rotundus L. (i.e., Nagar motha in India), Perilla frutescens L. (i.e., Basionym), and Aurantii Fructus Immaturus L., the natural flavanone glycosides Naringin and Neohesperidin.↓ CaSR-mediated NLRP3
inflammasome’s activation
Esophageal epithelial cells
(reflux esophagitis)
[333]
see also Box 1
XingxiongExtract from Ginkgo biloba L. or Ginkgo folium L. and tetramethylpyrazine sodium chloride ↓ NLRP3
↑Akt/NRF2 axis
Focal cerebral I/R damage[334]
Ze LanRhizomes or rootstalks of Lycopus lucidus↓ NLRP3 H2O2-induced oxidative injury
in rat embryo cortical neurons
[335]
§ ↑ = increased; ↓ = decreased.
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

Chiarini, A.; Gui, L.; Viviani, C.; Armato, U.; Dal Prà, I. NLRP3 Inflammasome’s Activation in Acute and Chronic Brain Diseases—An Update on Pathogenetic Mechanisms and Therapeutic Perspectives with Respect to Other Inflammasomes. Biomedicines 2023, 11, 999. https://doi.org/10.3390/biomedicines11040999

AMA Style

Chiarini A, Gui L, Viviani C, Armato U, Dal Prà I. NLRP3 Inflammasome’s Activation in Acute and Chronic Brain Diseases—An Update on Pathogenetic Mechanisms and Therapeutic Perspectives with Respect to Other Inflammasomes. Biomedicines. 2023; 11(4):999. https://doi.org/10.3390/biomedicines11040999

Chicago/Turabian Style

Chiarini, Anna, Li Gui, Chiara Viviani, Ubaldo Armato, and Ilaria Dal Prà. 2023. "NLRP3 Inflammasome’s Activation in Acute and Chronic Brain Diseases—An Update on Pathogenetic Mechanisms and Therapeutic Perspectives with Respect to Other Inflammasomes" Biomedicines 11, no. 4: 999. https://doi.org/10.3390/biomedicines11040999

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