Skip to main content

The NLRP3 inflammasome: contributions to inflammation-related diseases

Abstract

The NOD-like receptor protein 3 (NLRP3) inflammasome is a protein complex that regulates innate immune responses by activating caspase-1 and the inflammatory cytokines interleukin (IL)-1β and IL-18. Multiple studies have demonstrated the importance of the NLRP3 inflammasome in the development of immune and inflammation-related diseases, including arthritis, Alzheimer’s disease, inflammatory bowel disease, and other autoimmune and autoinflammatory diseases. This review first explains the activation and regulatory mechanism of the NLRP3 inflammasome. Secondly, we focus on the role of the NLRP3 inflammasome in various inflammation-related diseases. Finally, we look forward to new methods for targeting the NLRP3 inflammasome to treat inflammation-related diseases, and provide new ideas for clinical treatment.

Introduction

The innate immune system acts as the first line of host defense to trigger the adaptive immune response. This system initiates downstream inflammatory cascades in response to noxious stimuli through germline-encoded pattern recognition receptors (PRRs). PRRs are distributed in the cell membrane and cytoplasm, playing a prominent role in initiating innate and adaptive immunity. Their main function is to produce pro-inflammatory cytokines and interferons by transcription [1, 2]. The activating factors of PRRs include pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs), which are endogenous molecules derived from dying cells [3]. However, inappropriate activation of PRRs causes long-term inflammation and even leads to autoimmune and inflammatory diseases [4, 5]. As an important PRR in the cytoplasm, the NOD-like receptors (NLR) family is instrumental in the inflammatory response and has attracted wide attention in recent years.

Based on the nature of nitrogen-terminal domains, members of the NLRs family are subdivided into four subfamilies, including NLRA, NLRB, NLRC, and NLRP. NLRs family activates multiple downstream signals, promoting inflammasome assembly and inflammatory response [6]. In recent years, the NOD-like receptor protein 3 (NLRP3) and the inflammasome assembled from it is the most focused on inflammasome. However, the current understanding of the NLRP3 inflammasome still has unresolved questions, such as its structure and activation mechanism. In addition, increasing evidence suggests that activation of the NLRP3 inflammasome is involved in the pathological process of various inflammatory diseases [7,8,9]. The activation of the NLRP3 inflammasome aggravates oxidative stress and vascular endothelial dysfunction, and accelerates the pathological process of cardiovascular diseases [10]. In rheumatoid arthritis, the deletion of NLRP3 and its downstream components significantly reduces inflammation and cartilage destruction [11]. Furthermore, in the microglia of the nervous system, the NLRP3 inflammasome can sense protein misfolding deposition or amyloid β (Aβ) aggregation and be activated to promote the occurrence and progression of neurodegenerative diseases [12]. The important role of the NLRP3 inflammasome in human diseases further indicates that it is of great clinical significance to study the mechanism of its involvement in diseases and targeted drug therapy.

In this review, we first describe the composition of the NLRP3 inflammasome and elaborate on the activation mechanism of the NLRP3 inflammasome, including both canonical and non-canonical activation pathways. In addition, we highlight the pivotal role that the NLRP3 inflammasome plays in inflammation-related diseases. Finally, we look to the future of the NLRP3 inflammasome as a promising therapeutic target for disease.

NLRP3 inflammasome

The inflammasome is a crucial component of the innate immune system’s response to pathogens, which consists of a set of cytoplasmic multiprotein complexes. Inflammasomes activate specific caspase proteases in response to infection or noxious stimuli [13]. A variety of inflammasomes, including NLRP1, NLRP3, NLRC4, and AIM2, have been identified [6, 14]. Different inflammasomes have distinct stimulatory signals but have very conserved downstream effects, especially in the activation of caspase-1, which in turn triggers the three key substances: pro-interleukin-1β (IL-1β), pro-IL-18, and gasdermin D (GSDMD) [13, 15]. Most inflammasomes are activated only by one or a few highly specific agonists (e.g., AIM2 is activated only by DNA from DNA viruses or bacteria), whereas NLRP3 can respond to a variety of agonists that are abundant in source and unrelated in structure and chemical properties. As such, it possesses the broadest functional scope of all inflammasomes in both innate and adaptive immune systems [16, 17]. The NLRP3 protein belongs to the family of nucleotide-binding oligomerization domain-like receptors (NLRs) [18]. The NLRP3 protein contains a leucine-rich repeat (LRR) domain at the carboxyl terminus, a pyrin domain (PYD) at the amino terminus, and a nucleotide-binding domain (NACTH) in the center domain [18].

The inflammasome fails to assemble successfully without immune activators, which is due to an internal interaction between the NACHT domain and LRRs replacing the interaction between NLRP3 and the apoptosis-associated speck-like protein (ASC) [19]. However, when immune activators, such as PAMPs, DAMPs, and various exogenous or environmental stimuli are present, the internal structure of NLRP3 opens up, allowing PYD–PYD interaction between NLRP3 and ASC. Subsequently, the caspase recruitment domain (CARD) of ASC binds to the CARD on pro-caspase-1 to form the NLRP3 inflammasome. The inflammasome of NLRP3 produce mature caspase-1 that cleaves pro-IL-1β and pro-IL-18 to produce corresponding mature cytokines [20, 21].

Activation mechanism of the NLRP3 inflammasome

Canonical activation of NLRP3

The canonical activation of NLRP3 inflammasome is a two-step process consisting of priming and activation signals (Fig. 1).

Fig. 1
figure 1

Canonical activation of the NLRP3 inflammasome. The NLRP3 inflammasome comprises NLRP3 protein, ASC, and pro-caspase-1. The activation of the NLRP3 inflammasome includes both canonical and noncanonical methods. Multiple mechanisms activate the inflammasome, such as P2X7 channel-mediated K+ efflux, mitochondrial dysfunction, mitochondria active oxygen species (mtROS) release, lysosomal disruption, and subsequent cathepsin B release

Step 1: priming signal

Priming signals recognize DAMPs or PAMPs though Toll-like receptors and activate the NF-κB signaling pathway, upregulating IL-1β and NLRP3 protein expression. During priming, apoptosis-related signaling molecules caspase-8 and Fas-associated with death domain protein (FADD) act as regulators of NLRP3 inflammasome upstream signaling, affecting transcriptional initiation and post-translational activation of NLRP3 inflammasome pathways, independent of its apoptotic function [22]. Caspase-8 interacts with the inhibitory kappa B kinase (IKK) complex. The IKK complex is a fundamental component of activating the NF-κB pathway and promotes NF-κB transcription and translocation in initiation [23]. The priming signal has at least two functions for NLRP3 inflammasome activation. Firstly, up-regulated expression of related components, including NLRP3 protein, pro-caspase-1, and pro-IL-1β [24,25,26], and secondly to induce multiple posttranslational modifications (PTMs) of NLRP3, including ubiquitination, SUMOylation, and phosphorylation [27,28,29].

Step 2: activating signal

Various molecular or cellular events are important causes of NLRP3 inflammasome activation, including lipopolysaccharide (LPS), extracellular ATP, crystallization, ion flux, lysosomal damage, and production of reactive oxygen species (ROS) induced by mitochondrial dysfunction. Once activated, the inflammasome initiates a series of downstream reactions [30].

K+ efflux

Early studies have shown that K+ efflux is a common phenomenon that activates the NLRP3 inflammasome. NLRP3 agonists first trigger cell membrane permeation of K+ and Na+, which in turn activates the NLRP3 inflammasome. Notably, a low level of intracellular K+ concentration is sufficient to activate the NLRP3 inflammasome, whereas an increase in intracellular Na+ concentration is involved in regulation but is not necessary for NLRP3 inflammasome activation [31]. K+ efflux mediated by P2X7 receptor channels is an efficient pathway to activate the NLRP3 inflammasome [32]. The P2X7 receptor channel is a plasma membrane channel directly activated by extracellular ATP. ATP binds to the P2X7 receptor to form a DAMP and then opens the channel to cause K+ transmembrane efflux, ultimately activating the NLRP3 inflammasome. Similarly, in addition to the P2X7 receptor pathway, some two-pore domain potassium (K2P) channels can also cause the NLRP3 inflammasome activation in a K+ efflux-dependent manner. Fifteen K2P channel genes have been identified in the human genome, and six distinct subfamilies can be distinguished based on their structural and functional characteristics, including TASK, TALK, TRESK, TREK, THIK, and TWIK [33]. Among them, the ATP-induced two-pore domain weak inwardly rectifying K+ channel 2 (TWIK2) causes K+ efflux, leading to NLRP3 inflammasome activation in macrophages. It should be noted that there is a synergic effect between P2X7 and TWIK2, the former changes the membrane potential through Ca2+ and Na+ influx, while the latter activates the NLRP3 inflammasome through K+ efflux [34]. THIK-1, another member of the K2P channel family, is also involved in regulating NLRP3 inflammasome activation, and this process occurs downstream of the ATP/P2X7 receptor channel [35].

Furthermore, the mechanism of K+ efflux-mediated activation of the NLRP3 inflammasome involves the binding of never in mitosis gene A-related kinase 7 (NEK7) protein and NLRP3. NEK7 forms a molecular complex with NLRP3 in the cytoplasm, and its catalytic domain binds to the NLRP3 protein [36]. At present, the exact mechanism of K+ activation of the NLRP3 inflammasome is not completely clear, and no studies have shown that K+ directly interacts with the NLRP3 protein. Instead, it is believed that the NEK7–NLRP3 interaction may cause structural changes in the NLRP3 protein, resulting in K+-induced NLRP3 inflammasome activation [37]. To sum up, further exploration is required to fully understand the role of K+ efflux in activating the NLRP3 inflammasome.

Mitochondrial dysfunction and production of ROS

Mitochondrial dysfunction is also a key contributing factor for the NLRP3 inflammasome activation. The signals released by mitochondrial damage mainly include mitochondrial DNA (mtDNA) and mitochondria active oxygen species (mtROS). NLRP3 protein functions as a sensitive sensor, triggering downstream caspase-1 activation and releasing the cytokine IL-1β in response to cellular danger signals associated with mitochondrial damage [38, 39].

Shimada et al. first proposed that oxidized mitochondrial DNA (OX-mtDNA) could induce the activation of the NLRP3 inflammasome during apoptosis. In the presence of NF-κB, ATP induces mitochondrial dysfunction and cell apoptosis, leading to the release of OX-mtDNA from mitochondria. In the cytoplasm, OX-mtDNA binds and activates the NLRP3 inflammasome [40]. One study found that cytidine monophosphate kinase 2 (CMPK2) is involved in cytoplasmic OX-mtDNA-induced NLRP3 inflammasome activation. CMPK2 transcription provides deoxyribonucleotides for mtDNA synthesis, which lays the foundation for NLRP3 agonists to stimulate OX-mtDNA production [41]. Xue et al. demonstrated that increased levels of OX-mtDNA and aggravated mitophagy activate the NLRP3 inflammasome in rat liver [42]. The above results indicated that OX-mtDNA in the cytoplasm can successfully activate NLRP3 inflammasome.

In addition, mtROS not only directly regulates the assembly process of inflammasome, but also indirectly regulates the activity of inflammasome by affecting cytoplasmic proteins [43]. Specifically, mtROS mediates the dissociation of thioredoxin-interacting protein (TXNIP) from thioredoxin and then interacts with NLRP3 protein to activate the NLRP3 inflammasome [44]. The activation of the NF-κB pathway by ROS not only mediates the assembly of the NLRP3 inflammasome but also directly promotes the expression of tumor necrosis factor-α (TNF-α), pro-IL-1β, IL-6, and other inflammatory factors [45]. Another study showed that mitochondrial dysfunction activates the NLRP3 inflammasome through the synthesis of phosphatidylglycerol, but not ROS production. Phosphatidylglycerol directly binds to NLRP3 and interfering with its synthesis can specifically inhibit the activation of the NLRP3 inflammasome [46]. Furthermore, a recent study suggested that other organelles rather than mitochondria influence the activation of the NLRP3 inflammasome. After being stimulated, no morphological changes of NLRP3 puncta or colocalization of mitochondria with NLRP3 puncta are observed in mitochondria, especially NLRP3 puncta localized on the trans-golgi network (TGN). The recruitment of NLRP3 protein by the dispersive TGN as a scaffold leads to the polymerization of the adapter protein ASC, to activate downstream cascades [47]. The specific mechanisms of the NLRP3 inflammasome activation by mitochondrial dysfunction and resulting ROS production need to be further studied.

Lysosome rupture

An earlier study finds that lysosomes can uptake silica and aluminum salt in the alveoli, causing their swelling and rupture with concomitant exudation of the contents. This chain of events subsequently triggers activation of the NLRP3 inflammasome [48].

CD36 is a pattern recognition receptor that mediates the endocytosis of soluble ligands. CD36 are transformed into crystals or fibrils within cells, leading to lysosomal destruction and then the activation of NLRP3 inflammasome activates [49]. In a mouse model of Alzheimer’s disease (AD), the fibrillar peptide amyloid beta (Aβ) can be phagocytosed by lysosomes leading to lysosomal damage and cathepsin B release, which in turn activates the NLRP3 inflammasome [50]. Cathepsin B release not only directly activates the NLRP3 inflammasome but also mediates mitochondrial dysfunction, resulting in mtROS-triggered NLRP3 inflammasome activation [51]. Similarly, in a study of a mouse model of liver fibrosis, LPS/ATP treatment increases the contents of cathepsin B and ROS and activates NLRP3 inflammasome, while LPS/ATP and cathepsin B inhibitors fails to increase levels of NLRP3 and IL-1β [52].

Other activation signals

In addition to the three activation signals that have been previously described, there are several other signals that impact the initiation and activation of the NLRP3 inflammasome. For example, irrecoverable endoplasmic reticulum stress reduces the expression of microRNA 17 (miR17) through over-activated inositol-requiring enzyme 1. The reduction of miR17 leads to the stable expression of TXNIP protein, which activates the NLRP3 inflammasome and exacerbates apoptosis[53]. In addition to K+, other ion streams (including Na+, Ca2+, Mg2+, Mn2+, Zn2+, Fe2+, and Cl−) also participate in regulating the activation of NLRP3 inflammasome in macrophages [54].

Noncanonical activation of NLRP3 inflammasome

The noncanonical activation of the NLRP3 inflammasome is mainly mediated by lipopolysaccharide (LPS) on the surface of Gram-negative bacteria. Direct entry of LPS into the host cell cytoplasm occurs via endocytosis or transfection, without the requirement of Toll-like receptor 4 (TLR4) [55]. LPS binds and activates pro-caspase-11 upon entry into the cell, which further induces GSDMD cleavage. GSDMD belongs to the gasdermin family and has a functional N-terminal fragment and a self-inhibitory C-terminal fragment. Intracellular LPS binds and activates pro-caspase-11, this further induces GSDMD cleavage. The amino-terminal GSDMD fragment leads to pyroptosis and activates the NLRP3 inflammasome [56, 57]. Pyroptosis is a proinflammatory programmed cell death that results in the release of cellular contents. It is noteworthy that the noncanonical activation of the NLRP3 inflammasome in macrophages requires the cooperation of LPS and mRNA from Gram-negative bacteria. NLRP3 detects bacterial mRNA while LPS binds to pro-caspase-11. This interaction between pro-caspase-11 and NLRP3 enhances noncanonical activation of the NLRP3 inflammasome [58] (Fig. 2).

Fig. 2
figure 2

Noncanonical activation of the NLRP3 inflammasome. LPS can induce noncanonical activation of the NLRP3 inflammasome. LPS enters the cytoplasm by endocytosis or transfection, and then binds and activates pro-caspase-11, which subsequently cleaves GSDMD to promote pyroptosis, and activates the NLRP3 inflammasome

The activation of noncanonical NLRP3 inflammasome is dependent on GSDMD. The activation of GSDMD may not be potent enough to promote cell death, but it is robust enough to cause K+ efflux and activate NLRP3 inflammasome [59]. GSDMD not only alters the permeability of the cell membrane, but also creates pores in the mitochondrial membrane and causes the release of mtDNA. The presence of mtDNA and LPS allows Nur77 to bind to NLRP3 and activate NLRP3 inflammasome [60]. Additionally, once activated, the NLRP3 inflammasome promotes the cleavage of GSDMD by caspase-1 to produce N-GSDMD, which drives the onset of pyroptosis [61]. Pyroptosis caused by GSDMD leads to the release of IL-1β, promoting local inflammation [62]. GSDMD and NLRP3 reinforce each other, creating an amplified signaling loop. Precisely modulating this loop will provide new avenues for treating diseases.

The role of the NLRP3 inflammasome in inflammation-related diseases

NLRP3 inflammasome is widely involved in the pathological development of a variety of inflammation-related diseases. In this section, related diseases in several human vital organs/tissues are discussed (Table 1).

Table 1 The role of NLRP3 inflammasome in diseases

Autoinflammatory diseases

Cryopyrin-associated periodic syndrome

Cryopyrin-associated periodic syndrome (CAPS) is a group of rare genetic autoinflammatory diseases caused by mutations in the NLRP3 gene. CAPS can be classified into three subphenotypes according to the severity of the disease, including familial cold autoinflammatory syndrome (FCAS), Muckle–Wells syndrome (MWS), and neonatal multisystemic inflammatory syndrome (NOMID) (also known as chronic infantile neurocutaneous–articular syndrome, CINCA) [63]. In individuals with CAPS, gain-of-function mutations of the NLRP3 gene result in the excessive activation of NLRP3 inflammasome that cause sustained and uncontrolled release of IL-1β. [64]. The pathogenesis of CAPS may be related to the dysfunction of some negative regulators of NLRP3. For example, caspase recruitment domain-containing protein 8 (CARD8) can bind to NLRP3 to prevent exogenous stimulation. However, CAPS-associated NLRP3 mutations escape the control of CARD8 [65]. Furthermore, the cAMP–PKA directly inhibits the rapid activation of NLRP3 inflammasome, whereas CAPS-associated mutations disable this signaling pathway [66]. NLRP3 inflammasome activation-induced cytokine IL-1β is thought to be a key factor aggravating the inflammatory response [67] It is noteworthy that NLRP3 mutations appear to be specifically enhanced in neutrophils, with these cells being the primary source of IL-1β in severe CAPS, both in patients and in mouse models [68].

The current therapeutic approach for CAPS is to inhibit the downstream signaling pathway of IL-1, using the IL-1 receptor antagonist anakinra for clinical treatment [69, 70]. NLRP3 as an upstream component may be a potential new target. The current study found that the LRR sequence of NLRP3 plays an irreplaceable role in CAPS. Genes of the LRR sequence undergo alternative splicing after transcription, whose exons 4, 5, 7, and 9 determine whether NLRP3 inflammasome can be activated [71]. Importantly, this study shows that inhibition of LRR exons is sufficient to prevent the assembly of NLRP3 inflammasome, which could be a potential target for inhibitor development. Sensorineural hearing impairment is one of the common sufferings of CAPS patients. The NLRP3 inhibitor MCC950 significantly improved hearing impairment and systemic inflammation in CAPS mice [72]. In addition, the metalloproteinase inhibitor thiolutin prevented the deubiquitination of NLRP3 and thus the activation of multiple CAPS-associated NLRP3 inflammasome [27].

Familial Mediterranean fever

Familial Mediterranean fever (FMF) is a genetic disease caused by gene missense mutations. Mediterranean fever (MEFV) is the gene responsible for the pathogenesis of FMF and encodes the Pyrin protein. More than 50 mutations associated with FMF have been identified in MEFV [73, 74]. Pyrin can bind to ASC of multiple inflammasomes through its pyrin structural domain. In FMF, the pyrin inflammasome/caspase-1/GSDMD pathway causes the release of inflammatory cytokines and alarmins S100A8/A9 (hallmarks of FMF) extracellularly and exacerbates autoinflammation [75]. NLRP3 inflammasome trigger inflammation-related diseases mainly through the pro-inflammatory cytokine IL-1β. NLRP3 inflammasome appears to be more active in the population carrying MEFV mutations monocytes from FMF patients and healthy populations are treated with LPS, and monocytes from FMF patients secrete more IL-1β and positively correlate with the number and exocytosis of MEFV mutations. This IL-1β source is NLRP3-dependent, since silencing NLRP3 inhibits IL-1β secretion [76]. In addition, activation of the pyrin inflammasome in monocytes from FMF patients results in reduced expression of IL-1 receptor antagonists, making patients more sensitive to proinflammatory stimuli [77]. In contrast to the above studies, an animal study shows that the pathogenesis of FMF may not completely depend on the NLRP3 inflammasome, but rather be mediated by its downstream ASC or IL-1β agents[73]. The involvement of NLRP3 in the pathogenesis of FMF has not been fully elucidated. Furthermore, CAPS and FMF have similar pathogenesis and are both characterized by excessive IL-1 release, but their clinical manifestations are very different. Whether NLRP3 plays a key role in this is a potential research question.

Neurodegenerative diseases

Neurodegenerative diseases are characterized by the progressive neuronal loss [78]. The NLRP3 inflammasome is involved in and drives the development of neurodegenerative diseases [79, 80].

Alzheimer’s disease

AD is the main cause of dementia and is characterized by cognitive loss and memory impairment [81]. It is evaluated that the number of dementia patients in the world has now exceeded 50 million, with the annual cost of more than one trillion US dollars [82, 83]. AD has emerged as a pressing public health concern [82]. One of the classic pathological features of AD is an intracellular hyperphosphorylated tau protein composition of neurofibrillary tangles (NFTs) [84]. It has been demonstrated that the formation of NFT caused by NLRP3 inflammasome exacerbates the development of AD [85] The deletion of the NLRP3 inflammasome can reduce the hyperphosphorylation and aggregation of tau protein by regulating phosphatase and tau protein kinase, which improves cognitive deficit in mice [86].

Neuroinflammation is an important cause of NFTs formation and accelerates the pathological progression of AD [87]. Neuroinflammation in AD is mainly driven by microglia, and a genome-wide association study found that microglia express proteins encoded by mutated genes associated with late-onset AD [88]. Tau released by neurons into the extracellular matrix can activate the NLRP3 inflammasome in microglia, which causes the extracellular release of downstream IL-1β and IL-18. Persistent signaling of the NLRP3 inflammasome induces microglial dysfunction, weakens the clearance capability of NFTs by microglia, and ultimately sets up a vicious proinflammatory cycle, accelerating neuronal cell death [86].

Parkinson’s disease

Parkinson’s disease (PD) is a neurological disorder with rare occurrence, and its main clinical manifestations are characterized by resting tremor, slow movement, stiffness, and loss of postural reflex [89]. PD is a complex disorder with many etiological factors and is related to genetics, environment, and their interaction [90, 91]. The classic pathological mechanisms of PD mainly include the damage and apoptosis of dopaminergic neurons in the substantia nigra compacta and the accumulation of α-synuclein (α-Syn) [92, 93].

In recent years, there has been a significant focus on neuroinflammation and immune dysfunction in PD [94]. Activation of NLRP3 inflammasome in peripheral blood mononuclear cells of PD patients results in increased plasma levels of α-syn and IL-1β, both of which aggravate the severity of dyskinesia in PD patients. Interestingly, plasma α-Syn levels in PD patients are positively correlated with proinflammatory factor IL-1β levels, suggesting that α-Syn released by degenerated neuron may act as an endogenous substance to activate NLRP3 inflammasome and induce a strong inflammatory response in PD [95]. The interaction between aggregated α-Syn and Toll-like receptor 2 may serve as the first initiating signal of NLRP3 inflammasome activation. Internalization of α-Syn by microglia leads to overproduction of mtDNA and mtROS, which act as the second signal to the NLRP3 inflammasome activation and ultimately induce neuroinflammation [96, 97]. Meanwhile, the NLRP3 inflammasome can release IL-1β through caspase-1 cleavage and exacerbate the inflammatory response by inducing pyroptosis. The entry of IL-1β into the substantia nigra site can promote the loss of dopaminergic neurons [98].

NLRP3 inflammasome activation in microglia also plays a pivotal role in PD. In the neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-driven degeneration of dopaminergic neurons, the NLRP3 inflammasome activation in microglia significantly exacerbated dyskinesia and loss of dopaminergic neurons in mice [99]. In addition, microglial autophagy deficiency can result in the death of tyrosine hydrogenase neurons in the substantia nigra through NLRP3/PDE10A (phosphodiesterase 10A)–cyclic adenosine monophosphate (cAMP) signaling, the effects of which are reduced by NLRP3 inhibition [100]. The study by Qing et al. confirms this notion that autophagy deficiency in microglia can exacerbate MPTP-induced neurodegeneration in a mouse model by activating the NLRP3 inflammasome [101]

Drug therapy targeting the NLRP3 inflammasome improves the pathological progression of PD. Prussian blue nanozyme is a pyroptosis inhibitor, and can inhibit NLRP3 inflammasome activation by scavenging ROS as well as downregulate GSDMD autocleavage and pro-inflammatory factor production, attenuating MPTP neurotoxin induction neurodegeneration in mouse and cellular models of PD [102]. Additionally, dopamine inhibits canonical activation as well as noncanonical activation of the NLRP3 inflammasome in primary human microglia [103]. Andrographolide is a bicyclic diterpenoid lactone with immunomodulatory and antiinflammatory activities. Treatment of mice with andrographolide increases mitophagy, reduces activation of the NLRP3 inflammasome, and ultimately ameliorates the loss of dopaminergic neurons caused by the neurotoxin MPTP [104]. At present, the clinical drug therapy for PD still does not meet the ideal expectations. The animal experiments of antagonists targeting NLRP3 inflammasome and its related pathways have achieved favorable therapeutic effects. However, these drugs in the treatment of PD are intended for basic studies, and no clinical data have been evaluated.

Huntington’s disease

Huntington’s disease (HD) is an autosomal dominant neurodegenerative disease characterized by chorea, dystonia, motor incoordination, cognitive decline, and behavioral difficulties [105]. Galectin-3 (Gal-3) overexpression in microglia is one of the key causes of HD [106]. In a Sprague–Dawley rat model of spinal cord injury, Gal-3 aggravates neuroinflammation through the ROS/TXNIP/NLRP3 signaling pathway [107]. Similarly, Gal-3 expression is upregulated in the plasma of HD patients and mice. In vitro experiments further find that Gal-3 in the microglia of HD model produces IL-1β through the NLRP3 inflammasome-dependent pathway, aggravating brain inflammation [108].

Furthermore, studies in a transgenic R6/2 mouse model of HD (a model that exhibits a progressive neurological phenotype and mimics several features of human HD) find that NLRP3 and caspase-1 are highly elevated in the HD R6/2 mouse, which can further induce pyroptosis [109]. Notably, the NLRP3-specific inhibitor MCC950 treatment inhibits the activation of NLRP3 inflammasome, decreases IL-1β and ROS production, and ultimately improves motor dysfunction in mice [110]. Similarly, in another study, olaparib treatment downregulates the NLRP3 inflammasome and reduces caspase-1 cleavage in the R6/2 mouse model, ameliorating pyroptosis-induced missing neurons [111].

Cardiovascular diseases

A study on the association between NLRP3-related gene mutations and cardiovascular diseases mortality reports that NLRP3 intronic variant rs10754555 is associated with inflammasome activation and systemic inflammation, and carriers of this gene have a higher risk of death [112]. Low-grade basal NLRP3 inflammasome activation promotes the development of various chronic cardiovascular diseases such as hypertension and atherosclerosis [7, 113]. Consequently, inhibiting the activation of the NLRP3 inflammasome is crucial for the treatment of cardiovascular disease. This section mainly focuses on the NLRP3 inflammasome in cardiovascular disease.

Atherosclerosis is a common cardiovascular disease characterized by lipid accumulation and persistent inflammation in large and medium arteries, which can eventually lead to various complications such as myocardial infarction and cerebral infarction [114]. Recently, several studies have manifested that the NLRP3 inflammasome serves as a potential therapeutic target for atherosclerosis by regulating the expression of pro-inflammatory factors [113, 115]. Multiple endogenous danger signals, such as cholesterol crystals [116], oxidized low-density lipoprotein, and some endogenous metabolites [117] activate the NLRP3 inflammasome, enhancing the expression of proinflammatory cytokines. Ultimately, these danger signals contribute to the development of atherosclerosis by enhancing oxidative stress and the inflammatory response. In addition, multiple cellular dysfunctions can lead to overactivation of the NLRP3 inflammasome to aggravate atherosclerosis. Mitochondrial uncoupling protein 1 (UCP1) is a critical factor in the thermogenesis and mitochondrial function of brown fat cells. UCP1 depletion leads to hyperactivation of the IL-1β and NLRP3 inflammasome by increasing mitochondrial membrane potential and mitochondrial superoxide, thereby exacerbating endothelial dysfunction, vascular inflammation, and atherosclerosis in obese mice [115]. In addition, smoking and other bad living habits are also contributing factors of atherosclerosis. Specifically, nicotine treatment of human aortic endothelial cells results in NLRP3 inflammasome activation and pyroptosis, the effects of which are reversed by NLRP3 silencing or ASC with small interfering RNA [118].

Heart failure (HF) is a syndrome in which cardiac output cannot be maintained under normal filling pressure, mainly caused by a variety of cardiac structural or functional abnormalities [119]. Notably, activation of the NLRP3 inflammasome produces inflammatory factors that recruits macrophages and T cells to trigger cardiomyocyte pyroptosis and cardiac inflammation, leading to fibrosis, adverse cardiac remodeling, and even heart failure [120]. Zeng et al. found that in dilated cardiomyopathy (DCM) patients and doxorubicin-induced DCM mice, the NLRP3 inflammasome triggers cardiomyocyte pyroptosis through the caspase-1 pathway [121]. However, empagliflozin reduce cardiac inflammation in rats with heart failure and exert a cardioprotective effect by inhibiting the NLRP3 inflammasome [122].

Liver disease

Nonalcoholic steatohepatitis

Nonalcoholic fatty liver disease (NAFLD) is a clinicopathological syndrome that directly leads to cirrhosis and hepatocellular carcinoma [123, 124]. Nonalcoholic steatohepatitis (NASH) is one of the prime reasons of liver cirrhosis.

The foz/foz mouse is a model of over-nutrition with phenotypes of obesity and metabolic syndromes such as diabetes and hypercholesterolemia. Foz/foz mice fed a diet deficient in methionine and choline develop severe steatohepatitis and liver fibrosis. In the liver of foz/foz mice, cholesterol crystals promote the expression of NLRP3 and its downstream molecules caspase-1 and IL-1β. However, MCC950 treatment reduces the levels of these inflammatory factors and improves liver fibrosis [125]. Furthermore, in an obese mouse model fed a high-fat diet, the high-fat diet activates the hepatic NLRP3 inflammasome and enhances the expression of NLRP3, ASC, caspase-1, IL-6, and TNF-α, deteriorating NASH [126]. However, single-cell transcriptional profiling of the livers of Nlrp3A350V mutant mice show that activated NLRP3 can cause NASH and liver fibrosis even in the absence of hepatic steatosis [127]. Therefore, steatosis may not be necessary for the NLRP3 inflammasome-induced fibrosis and NASH.

MicroRNAs have a variety of functions in regulating gene expression of inflammatory factors [128, 129]. The specifically expressed microRNA-223 (miR-223) is significantly downregulated in neutrophils of the liver of NASH patients and is considered to be a key negative regulator of NLRP3 expression [130]. In miR-223-deficient mouse liver macrophages, the mRNA expression of NLRP3 is increased, while significantly decreased after supplementation of mice with miR-223 3p (miR-223 endogenous analog [131]). Furthermore, in fibrotic NASH mice chronically fed a high-fat, fructose, and cholesterol diet, miR-223 3p reduces NLRP3 and IL-1β production, significantly ameliorating NASH [132].

Liver cirrhosis

Cirrhosis is a common chronic liver disease, and its pathological changes include diffuse degeneration and necrosis of hepatocytes, fibrous tissue hyperplasia, and nodular regeneration of hepatocytes. Cirrhosis leads to the replacement of healthy liver tissue by fibrotic tissue and regenerative nodules, ultimately causing the loss of normal liver function [133]. Previous studies have found higher NLRP3 and caspase-1 expression levels in liver of patients with cirrhosis [134]. The NLRP3 inflammasome is significantly activated in a CCL4-induced mouse model of cirrhosis, which aggravates hepatocyte death and cirrhosis through the NLRP3/caspase-1/GSDMD classical pyroptosis pathway [135].

Progressive liver fibrosis is a common pathological process of liver cirrhosis [136], and is characterized by excessive deposition of extracellular matrix (ECM). Hepatic stellate cell (HSC) activation plays a main role in the pathogenesis of liver fibrosis [137]. The activated NLRP3/caspase-1 signaling pathway forms mature IL-18 and IL-1β, which further promotes the transformation of HSCs into myofibroblasts and generates ECM [138]. However, MCC950 can not only inhibit the development of fibrosis, but also improved the function of fibrotic liver. Auvro et al. show that MCC950 also inhibits collagen expression and HSC activation in the mouse model of fibrotic NASH [125]. Similarly, in the CCL4-induced liver fibrosis model, alpinetin and auranofin also exert antiinflammatory and antioxidant effects by blocking the NLRP3 inflammasome [139, 140].

The release of inflammasome components caused by hepatocyte pyroptosis is another pathological mechanism of liver fibrosis. Susanne et al. found that activation of the NLRP3 inflammasome could lead to hepatocyte pyroptosis in mouse and human primary hepatocytes, whereas it can be inhibited by caspase-1 inhibitor by blocking the activation of GSDMD. Components of the inflammasome that are released extracellularly activate HSCs through endocytosis, increase the secretion of IL-1ß and the expression of α-SMA, and promote the development of fibrosis [141]. The study by Alexander et al. also proves this view [142]. NLRP3 gene knockout in mice show liver cell pyroptosis, HSC activation, collagen deposition, severe neutrophil infiltration, and liver inflammation [142].

In addition, the abundance of Stenotrophomonas maltophilia is higher in the liver of patients with cirrhosis when compared with healthy controls. It is found that Stenotrophomonas maltophilia induces the formation of NLRP3 inflammasome complex by activating TLR4-mediated NF-κB signaling pathway, which drives cirrhosis to deteriorate to hepatocellular carcinoma in mice [143]. Together, these findings [142] suggest that the NLRP3 inflammasome activation plays an important role in the progression of liver fibrosis–cirrhosis–hepatocellular carcinoma.

Inflammatory bowel disease

Inflammatory bowel disease (IBD), incorporating ulcerative colitis (UC) and Crohn’s disease (CD), is characterized by intestinal inflammatory lesions. The NLRP3 inflammasome has dual roles in IBD. On the one hand, activation of the inflammasome further enhances the inflammatory response, leading to aggravation of colon damage; on the other hand, under certain conditions, the inflammasome can inhibit the further development of inflammation and protects the colon from further damage [144, 145].

Dextran sulfate sodium (DSS) is a common method for constructing colitis models. In DSS mice, the degree of colitis in NLRP3 knockout mice is milder than that in wild mice, partly due to the reduced levels of proinflammatory cytokines [144]. In addition, in the colonic mucosa of UC patients, the NLRP3 inflammasome is significantly upregulated, and its activity gradually increased with the degree of disease progression [146]. Further studies found that DSS could enhance the NLRP3 inflammasome activation in macrophages by increasing the K+ efflux mediated by calcium-activated intermediate-conductance potassium ion channel (KCa3.1). In addition, genetic studies targeting CD found that NLRP3-related single-nucleotide site mutations are associated with susceptibility to CD [147]. CARD8, a negative regulator of the NLRP3 inflammasome, prevents NLRP3 from binding to ASC during inflammasome activation. Mutation of the gene encoding the CARD8 protein can lead to enhanced NLRP3 activity and increase downstream IL-1β secretion, resulting in the occurrence of CD [148]. Notably, some NLRP3 nature original inhibitors can improve colitis by restraining the activation of NLRP3. For example, cardamonin, a natural flavone isolated from Alpinia katsumadai Hayata, attenuates mice colitis by activating the AhR/Nrf2/NQO1 pathway and inhibiting the activation of the NLRP3 inflammasome [149]. Ginsenoside RK3, the main active ingredient of ginseng, can alleviate DSS-induced ulcerative colitis by inhibiting the expression of the NLRP3 inflammasome [150].

Paradoxically, the NLRP3 inflammasome also has a certain protective role in intestinal inflammation. Mice with NLRP3, ASC, or caspase-1 gene knockout are more susceptible to DSS-induced colitis, which is a result of reduced IL-18, a downstream component of the NLRP3 inflammasome. Defective NLRP3 inflammasome activation results in loss of intestinal epithelial integrity, which in turn results in commensal bacterial overgrowth and bacteremia [145]. In addition, in the oxazolone-induced mouse UC model, NLRP3 knockout mice have more severe colitis, as indicated by increased Th2 cytokine expression and decreased production of mature IL-1β and IL-18. Exogenous administration of IL-1β reduces the expression of colonic Th2 cytokines IL-4 and IL-13 and improves colitis, while exogenous IL-18 also reduces the severity of colitis but does not affect the expression of Th2 cytokines [151]. Yao et al. found that NLRP3R258W mutant mice have a gain-of-function mutation in the NLRP3 gene coding region. Notably, the mice suffer from CAPS but do not develop autoinflammation because the NLRP3R258W mutation-mediated remodeling of intestinal microbiota induces Treg cells to maintain homeostasis and improve intestinal anti-inflammatory ability [152].

Arthritis

Rheumatoid arthritis

Rheumatoid arthritis (RA) is an inflammatory disease that primarily affects joints throughout the body, and can lead to cartilage and bone damage and even disability [153]. Many inflammatory factors such as TNF-α, IL-1β, and IL-6 are involved in the occurrence and development of inflammatory joint damage [154]. Genetic studies have shown that the C allele at rs4612666 and the G allele at rs10754558 of the NLRP3 gene coding locus can increase the risk of RA [155]. The NLRP3 inflammasome is mainly involved in RA progression through regulating downstream cytokines, of which inhibition of IL-1β is particularly significant for RA treatment [83]. The NLRP3 inflammasome in synovial fluid monocytes and macrophages is significantly activated in RA patients and in a mouse model of collagen-induced arthritis. MCC950 treatment significantly reduces IL-1β and improves joint inflammation and bone destruction in a mouse model [156]. In addition, RA disease itself may also trigger the activation of the NLRP3 inflammasome, forming a vicious circle and aggravating the development of the disease. However, anticitrullinated protein antibodies (ACPAs) (RA-specific autoantibodies) can induce the expression of NLRP3 and pro-IL-1β by activating the CD147/ITGB1/Akt/NF-κB signaling pathway. On the other hand, ACPAs can activate Pannexin channels to promote the release of ATP. Subsequently, the accumulated ATP binds to the P2X7 receptor leading to the activation of NLRP3 inflammasome [157]. Moreover, extracellular Ca2+ promotes calcium-sensing receptor signaling in RA, leading to the activation of the NLRP3 inflammasome and the release of IL-1β [158].

Gout

Gout is a recurrent inflammatory disease caused by the deposition of urate crystals in the synovium, synovial bursa, cartilage, and other tissues of joints. Activation of NLRP3 inflammasome and release of IL-1β by monosodium urate crystals are pivotal pathological factors in gout attacks [159]. Fabio et al. revealed that monosodium urate (MSU) is involved in the activation of the NLRP3 inflammasome and promotes the production of IL-1β and IL-18 [160]. Some human experiments have shown that IL-1β inhibitors rilonacept, canakinumab, and anakinra are effective in the treatment of acute and chronic gout patients [161, 162]. In addition, a ketogenic diet (KD) can inhibit NLRP3 inflammasome activation by increasing the level of beta-hydroxybutyrate in neutrophils, thereby blocking IL-1β production and alleviating gout [163]. In conclusion, the NLRP3 inflammasome is expected to be a future therapeutic target for gout.

Summary and perspectives

The innate immune system provides the host with the first line of defense against pathogens, but continued activation of this system can lead to several diseases. The NLRP3 inflammasome is a key component of the innate immune system and plays an important role in inflammation-related diseases such as atherosclerosis, AD, and IBD through promoting the release of IL-1β and IL-18. This review mainly summarized the various activating factors of NLRP3 inflammasome (Fig. 1) and emphasizes its regulatory mechanism in the above-mentioned diseases, providing new ideas for targeting NLRP3 inflammasome to treat inflammation-related diseases. The activation mechanism of NLRP3 inflammasome mainly includes K+ efflux, mitochondrial dysfunction, and lysosome rupture. However, most of the current understanding of the pathogenic mechanism of NLRP3 inflammasome comes from animal experiments, and how it is activated and regulated in humans has not been fully elucidated. Therefore, the study of NLRP3 inflammasome compositions, downstream pathways, tandem pathways, and the development of related inhibitors have broad prospects in the treatment of a wide range of inflammation-related diseases.

A variety of NLRP3 inhibitors have been found to target the NLRP3 inflammasome, some of which directly target the NLRP3 protein, while others target other components and downstream products of NLRP3 inflammasome. For example, the monoclonal anti-IL-1β antibody canakinumab can rapidly relieve the symptoms of CAPS patients and has a high safety profile [164, 165]. In addition, NLRP3-specific small-molecule inhibitor MCC950 has shown promising efficacy in various animal models of different diseases, such as autoinflammatory diseases, cardiovascular diseases, cancer, neurological diseases, and diabetes[166,167,168,169]. MCC950 can inhibit ATP hydrolysis and NLRP3 inflammasome formation by interacting with the Walker B motif in the NACHT domain of NLRP3 protein [170,171,172]. Another endogenous small molecule inhibitor BHB reduces K+ efflux and ASC speck-like aggregates formation to inhibit NLRP3 inflammasome activation, decreasing the levels of IL-1β and IL-18 produced by NLRP3 inflammasome in human monocytes [173]. BHB has been confirmed to play an anti-inflammatory role in the animal models of AD, gout, and acute kidney injury, and is expected to become a potential clinical drug [163, 174, 175]. However, the clinical application of NLRP3 inhibitors also has a lot of limitations. Previous studies have shown that the pharmacokinetics and toxicokinetics of MCC950 limit its clinical application [176]. Although BHB can prevent chronic progressive diseases such as diabetes and AD, BHB has a poor therapeutic effect in the face of acute inflammatory diseases. Improving the defects of existing NLRP3 inflammasome inhibitors, such as reducing the hepatotoxicity of MCC950, or developing new NLRP3 inflammasome targeting inhibitors is an urgent problem to be solved, which may provide new idea for solving clinically related diseases.

In addition, noncoding RNAs, such as microRNAs, also provide new possibilities for inhibiting the NLRP3 inflammasome [131]. LncRNA 4344 directly reduces the expression of NLRP3 and its downstream genes by targeting miR-138-5p, and improves cognitive behavior, neuroinflammation, and apoptosis in rats with LPS-induced cognitive impairment [177]. Long noncoding RNA XLOC_000647 can inhibit NLRP3 promoter activity and reduce NLRP3 expression at the transcriptional level in mice with pancreatic cancer, subsequently retarding cancer cells proliferation, invasion, and the epithelial–mesenchymal transition of in vitro [178]. Noncoding RNAs can regulate the expression and function of NLRP3 inflammasomes more finely at the gene level. Therefore, an in-depth understanding of its mechanism may help develop more precise and effective targets for NLRP3 inhibition.

Notably, targeting NEK7 provides a new direction for the development of NLRP3 inhibitors. NEK7 is a serine/threonine protein kinase that helps spindle formation and drives mitosis [179]. The formation of the NLRP3–NEK7 complex is one of the critical steps in the activation of the NLRP3 inflammasome [36, 37]. Previous studies have shown that artemisinin improves joint swelling in arthritic mice through inhibiting the NLRP3–NEK7 interaction and attenuating LPS- and MSU-induced K+ efflux in macrophages [180]. Novelly, RRx-001, a well-tolerated anticancer drug, can also block NLRP3–NEK7 interaction, alleviating colitis induced by DSS, systemic inflammation induced by LPS, and experimental autoimmune encephalomyelitis in mice [181]. Overall, most of the studies on NLRP3 targeting inhibitors are still at the stage of animal experiments. There are some NLRP3 inhibitors have entered early clinical trials, but the best clinical indications for NLRP3 inflammasome blockade remain to be determined [182]. The broad role of NLRP3 in disease poses a great challenge in selecting the indication for the best efficacy.

Interestingly, the NLRP3 inflammasome has two sides in inflammation-related diseases. The NLRP3 inflammasome exhibits deleterious effects in most diseases, including inflammatory diseases, metabolic diseases, and most tumors [183]. However, in some cancers, infectious diseases, and IBD, the NLRP3 inflammasome has different roles [151, 183]. For example, the NLRP3/IL-1 signaling axis induces the expansion of bone marrow-derived suppressor cells in melanoma, which reduces the activity of natural killer cell and CD8+ T cell activity and accelerates tumor growth [184]. However, lack of the NLRP3 inflammasome results in impaired IL-18 signaling, which in turn leads to liver metastasis of colorectal cancer in mice with colorectal cancer [185]. However, the reasons for the above differences are still unknown, which may be related to the involvement of NLRP3 inflammasome in different signaling pathways. How to precisely regulate the activation and function of the NLRP3 inflammasome so that it can play a more protective role in different organs and diseases may be a new direction for future research.

In conclusion, the research on NLRP3 and inflammation-related diseases has made some progress, but there are still many challenges, such as the precise regulation of NLRP3 inflammasome activation and function, and the development and clinical application of NLRP3 targeted inhibitors. However, the complicated interaction of different molecular pathways involved in NLRP3 activation increases the difficulty of targeted therapy. In the meantime, elucidating the mechanisms by which NLRP3 inflammasome plays dual roles in some diseases, such as IBD and colorectal cancers, will promote the research of inflammation-related diseases in the future. Therefore, based on the previous researches, a lot of basic and clinical researches are still needed to achieve targeting the NLRP3 inflammasome in the treatment of inflammation-related diseases.

Availability of data and materials

Not applicable.

Abbreviations

Aβ:

Amyloid β

ACPAs:

Anticitrullinated protein antibodies

AD:

Alzheimer’s disease

ASC:

Apoptosis-associated speck-like protein

α-Syn:

α-Synuclein

cAMP:

Cyclic adenosine monophosphate

CAPS:

Cryopyrin-associated periodic syndrome

CARD:

Caspase recruitment domain

CD:

Crohn’s disease

CINCA:

Chronic infantile neurologic, cutaneous, and articular

CMPK2:

Cytidine monophosphate kinase 2

CNKI:

China National Knowledge Infrastructure

DAMPs:

Damage-associated molecular patterns

DCM:

Dilated cardiomyopathy

DSS:

Dextran sulfate sodium

ECM:

Extracellular matrix

FADD:

Fas-associated with death domain protein

FCAS:

Familial cold autoinflammatory syndrome

FMF:

Familial Mediterranean fever

Gal-3:

Galectin-3

GSDMD:

Gasdermin D

HD:

Huntington’s disease

HF:

Heart failure

HSC:

Hepatic stellate cell

IBD:

Inflammatory bowel disease

IL-1β:

Interleukin-1β

IKK:

Inhibitory kappa B kinase

KD:

Ketogenic diet

K2P:

Two-pore domain potassium

LPS:

Lipopolysaccharide

LRR:

Leucine-rich repeat

NACTH:

Nucleoside triphosphatase domain

NAFLD:

Nonalcoholic fatty liver disease

NASH:

Nonalcoholic steatohepatitis

NEK7:

NIMA (never in mitosis gene A)-related kinase 7

NFTs:

Neurofibrillary tangles

NLR:

NOD-like receptors

NLRP3:

NOD-like receptor protein 3

NOMID:

Neonatal-onset multisystem inflammatory disorder

MEFV :

Mediterranean fever

miR:

MicroRNA

MPTP:

1-Methyl-4-phenyl-1,2,3,6-tetrahydropyridine

MSU:

Monosodium urate

MWS:

Muckle–Wells syndrome

mtDNA:

Mitochondrial DNA

mtROS:

Mitochondria active oxygen species

OX-mtDNA:

Oxidized mitochondrial DNA

PAMPs:

Pathogen-associated molecular patterns

PD:

Parkinson’s disease

PDE10A:

Phosphodiesterase 10A

PRRs:

Pattern recognition receptors

PTMs:

Posttranslational modifications

PYD:

Pyrimidine domain

RA:

Rheumatoid arthritis

ROS:

Reactive oxygen species

TGN:

Trans-Golgi network

TLR4:

Toll-like receptor 4

TNF-α:

Tumor necrosis factor-α

TWIK2:

Two-pore domain weak inwardly rectifying K+ channel 2

TXNIP:

Thioredoxin-interacting protein

UC:

Ulcerative colitis

UCP1:

Uncoupling protein 1

References

  1. Panda S, Nilsson JA, Gekara NO. Deubiquitinase MYSM1 regulates innate immunity through inactivation of TRAF3 and TRAF6 complexes. Immunity. 2015;43(4):647–59.

    Article  CAS  PubMed  Google Scholar 

  2. Sun R, Hedl M, Abraham C. IL23 induces IL23R recycling and amplifies innate receptor-induced signalling and cytokines in human macrophages, and the IBD-protective IL23R R381Q variant modulates these outcomes. Gut. 2020;69(2):264–73.

    Article  CAS  PubMed  Google Scholar 

  3. Zhang Q, Raoof M, Chen Y, Sumi Y, Sursal T, Junger W, et al. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature. 2010;464(7285):104–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Chen K, Wu T, Wang D, Li R, Shen X, Zhao T, et al. Transcriptomics and quantitative proteomics reveal changes after second stimulation of bone marrow-derived macrophages from lupus-prone MRL/lpr mice. Front Immunol. 2022;13:1004232.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Ranjan K, Hedl M, Sinha S, Zhang X, Abraham C. Ubiquitination of ATF6 by disease-associated RNF186 promotes the innate receptor-induced unfolded protein response. J Clin Invest. 2021. https://doi.org/10.1172/JCI145472.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Chou WC, Jha S, Linhoff MW, Ting JP. The NLR gene family: from discovery to present day. Nat Rev Immunol. 2023. https://doi.org/10.1038/s41577-023-00849-x.

    Article  PubMed  Google Scholar 

  7. Higashikuni Y, Liu W, Numata G, Tanaka K, Fukuda D, Tanaka Y, et al. NLRP3 inflammasome activation through heart-brain interaction initiates cardiac inflammation and hypertrophy during pressure overload. Circulation. 2023;147(4):338–55.

    Article  CAS  PubMed  Google Scholar 

  8. Shahzad K, Fatima S, Khawaja H, Elwakiel A, Gadi I, Ambreen S, et al. Podocyte-specific Nlrp3 inflammasome activation promotes diabetic kidney disease. Kidney Int. 2022;102(4):766–79.

    Article  CAS  PubMed  Google Scholar 

  9. de Carvalho Ribeiro M, Iracheta-Vellve A, Babuta M, Calenda CD, Copeland C, Zhuang Y, et al. Alcohol-induced extracellular asc specks perpetuate liver inflammation and damage in alcoholic hepatitis even after alcohol cessation. Hepatology. 2023. https://doi.org/10.1097/HEP.0000000000000298.

    Article  PubMed  Google Scholar 

  10. Ma P, Zha S, Shen X, Zhao Y, Li L, Yang L, et al. NFAT5 mediates hypertonic stress-induced atherosclerosis via activating NLRP3 inflammasome in endothelium. Cell Commun Signal. 2019;17(1):102.

    Article  PubMed  PubMed Central  Google Scholar 

  11. VandeWalle L, Van Opdenbosch N, Jacques P, Fossoul A, Verheugen E, Vogel P, et al. Negative regulation of the NLRP3 inflammasome by A20 protects against arthritis. Nature. 2014;512(7512):69–73.

    Article  Google Scholar 

  12. Heneka MT, McManus RM, Latz E. Inflammasome signalling in brain function and neurodegenerative disease. Nat Rev Neurosci. 2018;19(10):610–21.

    Article  CAS  PubMed  Google Scholar 

  13. Wang L, Fu H, Nanayakkara G, Li Y, Shao Y, Johnson C, et al. Novel extracellular and nuclear caspase-1 and inflammasomes propagate inflammation and regulate gene expression: a comprehensive database mining study. J Hematol Oncol. 2016;9(1):122.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Junqueira C, Crespo A, Ranjbar S, de Lacerda LB, Lewandrowski M, Ingber J, et al. FcgammaR-mediated SARS-CoV-2 infection of monocytes activates inflammation. Nature. 2022;606(7914):576–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Prochnicki T, Vasconcelos MB, Robinson KS, Mangan MSJ, De Graaf D, Shkarina K, et al. Mitochondrial damage activates the NLRP10 inflammasome. Nat Immunol. 2023;24(4):595–603.

    Article  CAS  PubMed  Google Scholar 

  16. Lamkanfi M, Dixit VM. Inflammasomes and their roles in health and disease. Annu Rev Cell Dev Biol. 2012;28:137–61.

    Article  CAS  PubMed  Google Scholar 

  17. Zhang Y, Yang W, Li W, Zhao Y. NLRP3 inflammasome: checkpoint connecting innate and adaptive immunity in autoimmune diseases. Front Immunol. 2021;12: 732933.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Xiao L, Magupalli VG, Wu H. Cryo-EM structures of the active NLRP3 inflammasome disc. Nature. 2023;613(7944):595–600.

    Article  CAS  PubMed  Google Scholar 

  19. Inoue M, Shinohara ML. NLRP3 inflammasome and MS/EAE. Autoimmune Dis. 2013;2013: 859145.

    PubMed  PubMed Central  Google Scholar 

  20. Lu A, Magupalli VG, Ruan J, Yin Q, Atianand MK, Vos MR, et al. Unified polymerization mechanism for the assembly of ASC-dependent inflammasomes. Cell. 2014;156(6):1193–206.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Brydges SD, Broderick L, McGeough MD, Pena CA, Mueller JL, Hoffman HM. Divergence of IL-1, IL-18, and cell death in NLRP3 inflammasomopathies. J Clin Invest. 2013;123(11):4695–705.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Gurung P, Anand PK, Malireddi RK, VandeWalle L, Van Opdenbosch N, Dillon CP, et al. FADD and caspase-8 mediate priming and activation of the canonical and noncanonical Nlrp3 inflammasomes. J Immunol. 2014;192(4):1835–46.

    Article  CAS  PubMed  Google Scholar 

  23. Su H, Bidere N, Zheng L, Cubre A, Sakai K, Dale J, et al. Requirement for caspase-8 in NF-kappaB activation by antigen receptor. Science. 2005;307(5714):1465–8.

    Article  CAS  PubMed  Google Scholar 

  24. Li S, Fang Y, Zhang Y, Song M, Zhang X, Ding X, et al. Microglial NLRP3 inflammasome activates neurotoxic astrocytes in depression-like mice. Cell Rep. 2022;41(4): 111532.

    Article  CAS  PubMed  Google Scholar 

  25. Xu Z, Zha X, Ji R, Zhao H, Zhou S. Green biosynthesis of silver nanoparticles using aqueous extracts of ageratum conyzoides and their anti-inflammatory effects. ACS Appl Mater Interfaces. 2023. https://doi.org/10.1021/acsami.2c22114.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Xie CB, Qin L, Li G, Fang C, Kirkiles-Smith NC, Tellides G, et al. Complement membrane attack complexes assemble NLRP3 inflammasomes triggering IL-1 activation of IFN-gamma-primed human endothelium. Circ Res. 2019;124(12):1747–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Ren GM, Li J, Zhang XC, Wang Y, Xiao Y, Zhang XY, et al. Pharmacological targeting of NLRP3 deubiquitination for treatment of NLRP3-associated inflammatory diseases. Sci Immunol. 2021. https://doi.org/10.1126/sciimmunol.abe2933.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Spalinger MR, Kasper S, Gottier C, Lang S, Atrott K, Vavricka SR, et al. NLRP3 tyrosine phosphorylation is controlled by protein tyrosine phosphatase PTPN22. J Clin Invest. 2016;126(5):1783–800.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Qin Y, Li Q, Liang W, Yan R, Tong L, Jia M, et al. TRIM28 SUMOylates and stabilizes NLRP3 to facilitate inflammasome activation. Nat Commun. 2021;12(1):4794.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Ohto U, Kamitsukasa Y, Ishida H, Zhang Z, Murakami K, Hirama C, et al. Structural basis for the oligomerization-mediated regulation of NLRP3 inflammasome activation. Proc Natl Acad Sci U S A. 2022;119(11): e2121353119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Munoz-Planillo R, Kuffa P, Martinez-Colon G, Smith BL, Rajendiran TM, Nunez G. K(+) efflux is the common trigger of NLRP3 inflammasome activation by bacterial toxins and particulate matter. Immunity. 2013;38(6):1142–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Karmakar M, Katsnelson MA, Dubyak GR, Pearlman E. Neutrophil P2X7 receptors mediate NLRP3 inflammasome-dependent IL-1beta secretion in response to ATP. Nat Commun. 2016;7:10555.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Piechotta PL, Rapedius M, Stansfeld PJ, Bollepalli MK, Ehrlich G, Andres-Enguix I, et al. The pore structure and gating mechanism of K2P channels. EMBO J. 2011;30(17):3607–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Di A, Xiong S, Ye Z, Malireddi RKS, Kometani S, Zhong M, et al. The TWIK2 potassium efflux channel in macrophages mediates NLRP3 inflammasome-induced inflammation. Immunity. 2018;49(1):56-65 e4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Drinkall S, Lawrence CB, Ossola B, Russell S, Bender C, Brice NB, et al. The two pore potassium channel THIK-1 regulates NLRP3 inflammasome activation. Glia. 2022;70(7):1301–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. He Y, Zeng MY, Yang D, Motro B, Nunez G. NEK7 is an essential mediator of NLRP3 activation downstream of potassium efflux. Nature. 2016;530(7590):354–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Sharif H, Wang L, Wang WL, Magupalli VG, Andreeva L, Qiao Q, et al. Structural mechanism for NEK7-licensed activation of NLRP3 inflammasome. Nature. 2019;570(7761):338–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zhou R, Yazdi AS, Menu P, Tschopp J. A role for mitochondria in NLRP3 inflammasome activation. Nature. 2011;469(7329):221–5.

    Article  CAS  PubMed  Google Scholar 

  39. Wu X, Gong L, Xie L, Gu W, Wang X, Liu Z, et al. NLRP3 deficiency protects against intermittent hypoxia-induced neuroinflammation and mitochondrial ROS by promoting the PINK1-Parkin pathway of mitophagy in a murine model of sleep apnea. Front Immunol. 2021;12: 628168.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Shimada K, Crother TR, Karlin J, Dagvadorj J, Chiba N, Chen S, et al. Oxidized mitochondrial DNA activates the NLRP3 inflammasome during apoptosis. Immunity. 2012;36(3):401–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zhong Z, Liang S, Sanchez-Lopez E, He F, Shalapour S, Lin XJ, et al. New mitochondrial DNA synthesis enables NLRP3 inflammasome activation. Nature. 2018;560(7717):198–203.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Jia X, Qiu T, Yao X, Jiang L, Wang N, Wei S, et al. Arsenic induces hepatic insulin resistance via mtROS-NLRP3 inflammasome pathway. J Hazard Mater. 2020;399: 123034.

    Article  CAS  PubMed  Google Scholar 

  43. Liang H, Zhang F, Wang W, Zhao W, Zhou J, Feng Y, et al. Heat shock transcription factor 2 promotes mitophagy of intestinal epithelial cells through PARL/PINK1/Parkin pathway in ulcerative colitis. Front Pharmacol. 2022;13: 893426.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Zhou R, Tardivel A, Thorens B, Choi I, Tschopp J. Thioredoxin-interacting protein links oxidative stress to inflammasome activation. Nat Immunol. 2010;11(2):136–40.

    Article  CAS  PubMed  Google Scholar 

  45. Pang Y, Wu D, Ma Y, Cao Y, Liu Q, Tang M, et al. Reactive oxygen species trigger NF-kappaB-mediated NLRP3 inflammasome activation involvement in low-dose CdTe QDs exposure-induced hepatotoxicity. Redox Biol. 2021;47: 102157.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Iyer SS, He Q, Janczy JR, Elliott EI, Zhong Z, Olivier AK, et al. Mitochondrial cardiolipin is required for Nlrp3 inflammasome activation. Immunity. 2013;39(2):311–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen J, Chen ZJ. PtdIns4P on dispersed trans-Golgi network mediates NLRP3 inflammasome activation. Nature. 2018;564(7734):71–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Hornung V, Bauernfeind F, Halle A, Samstad EO, Kono H, Rock KL, et al. Silica crystals and aluminum salts activate the NALP3 inflammasome through phagosomal destabilization. Nat Immunol. 2008;9(8):847–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sheedy FJ, Grebe A, Rayner KJ, Kalantari P, Ramkhelawon B, Carpenter SB, et al. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat Immunol. 2013;14(8):812–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Halle A, Hornung V, Petzold GC, Stewart CR, Monks BG, Reinheckel T, et al. The NALP3 inflammasome is involved in the innate immune response to amyloid-beta. Nat Immunol. 2008;9(8):857–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Hofbauer D, Mougiakakos D, Broggini L, Zaiss M, Buttner-Herold M, Bach C, et al. beta2-microglobulin triggers NLRP3 inflammasome activation in tumor-associated macrophages to promote multiple myeloma progression. Immunity. 2021;54(8):1772-87 e9.

    Article  CAS  PubMed  Google Scholar 

  52. Inzaugarat ME, Johnson CD, Holtmann TM, McGeough MD, Trautwein C, Papouchado BG, et al. NLR family pyrin domain-containing 3 inflammasome activation in hepatic stellate cells induces liver fibrosis in mice. Hepatology. 2019;69(2):845–59.

    Article  CAS  PubMed  Google Scholar 

  53. Lerner AG, Upton JP, Praveen PV, Ghosh R, Nakagawa Y, Igbaria A, et al. IRE1alpha induces thioredoxin-interacting protein to activate the NLRP3 inflammasome and promote programmed cell death under irremediable ER stress. Cell Metab. 2012;16(2):250–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Li C, Chen M, He X, Ouyang D. A mini-review on ion fluxes that regulate NLRP3 inflammasome activation. Acta Biochim Biophys Sin (Shanghai). 2021;53(2):131–9.

    Article  CAS  PubMed  Google Scholar 

  55. Kayagaki N, Wong MT, Stowe IB, Ramani SR, Gonzalez LC, Akashi-Takamura S, et al. Noncanonical inflammasome activation by intracellular LPS independent of TLR4. Science. 2013;341(6151):1246–9.

    Article  CAS  PubMed  Google Scholar 

  56. Kayagaki N, Warming S, Lamkanfi M, VandeWalle L, Louie S, Dong J, et al. Non-canonical inflammasome activation targets caspase-11. Nature. 2011;479(7371):117–21.

    Article  CAS  PubMed  Google Scholar 

  57. Kayagaki N, Stowe IB, Lee BL, O’Rourke K, Anderson K, Warming S, et al. Caspase-11 cleaves gasdermin D for non-canonical inflammasome signalling. Nature. 2015;526(7575):666–71.

    Article  CAS  PubMed  Google Scholar 

  58. Moretti J, Jia B, Hutchins Z, Roy S, Yip H, Wu J, et al. Caspase-11 interaction with NLRP3 potentiates the noncanonical activation of the NLRP3 inflammasome. Nat Immunol. 2022;23(5):705–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. de Sa KSG, Amaral LA, Rodrigues TS, Ishimoto AY, de Andrade WAC, de Almeida L, et al. Gasdermin-D activation promotes NLRP3 activation and host resistance to Leishmania infection. Nat Commun. 2023;14(1):1049.

    Article  PubMed  PubMed Central  Google Scholar 

  60. Zhu F, Ma J, Li W, Liu Q, Qin X, Qian Y, et al. The orphan receptor Nur77 binds cytoplasmic LPS to activate the non-canonical NLRP3 inflammasome. Immunity. 2023;56(4):753-67 e8.

    Article  CAS  PubMed  Google Scholar 

  61. Chang M, Wang Z, Dong C, Zhou R, Chen L, Huang H, et al. Ultrasound-amplified enzyodynamic tumor therapy by perovskite nanoenzyme-enabled cell pyroptosis and cascade catalysis. Adv Mater. 2023;35(7): e2208817.

    Article  PubMed  Google Scholar 

  62. Sun L, Li X, Luo Z, Li M, Liu H, Zhu Z, et al. Purinergic receptor P2X7 contributes to abdominal aortic aneurysm development via modulating macrophage pyroptosis and inflammation. Transl Res. 2023. https://doi.org/10.1016/j.trsl.2023.03.002.

    Article  PubMed  Google Scholar 

  63. Nakagawa K, Gonzalez-Roca E, Souto A, Kawai T, Umebayashi H, Campistol JM, et al. Somatic NLRP3 mosaicism in Muckle-Wells syndrome. A genetic mechanism shared by different phenotypes of cryopyrin-associated periodic syndromes. Ann Rheum Dis. 2015;74(3):603–10.

    Article  CAS  PubMed  Google Scholar 

  64. Bertoni A, Carta S, Baldovini C, Penco F, Balza E, Borghini S, et al. A novel knock-in mouse model of cryopyrin-associated periodic syndromes with development of amyloidosis: therapeutic efficacy of proton pump inhibitors. J Allergy Clin Immunol. 2020;145(1):368-78 e13.

    Article  CAS  PubMed  Google Scholar 

  65. Ito S, Hara Y, Kubota T. CARD8 is a negative regulator for NLRP3 inflammasome, but mutant NLRP3 in cryopyrin-associated periodic syndromes escapes the restriction. Arthritis Res Ther. 2014;16(1):R52.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Mortimer L, Moreau F, MacDonald JA, Chadee K. NLRP3 inflammasome inhibition is disrupted in a group of auto-inflammatory disease CAPS mutations. Nat Immunol. 2016;17(10):1176–86.

    Article  CAS  PubMed  Google Scholar 

  67. Lachmann HJ, Lowe P, Felix SD, Rordorf C, Leslie K, Madhoo S, et al. In vivo regulation of interleukin 1beta in patients with cryopyrin-associated periodic syndromes. J Exp Med. 2009;206(5):1029–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Stackowicz J, Gaudenzio N, Serhan N, Conde E, Godon O, Marichal T, et al. Neutrophil-specific gain-of-function mutations in Nlrp3 promote development of cryopyrin-associated periodic syndrome. J Exp Med. 2021. https://doi.org/10.1084/jem.20201466.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Romano M, Arici ZS, Piskin D, Alehashemi S, Aletaha D, Barron KS, et al. The 2021 EULAR/American College of Rheumatology points to consider for diagnosis, management and monitoring of the interleukin-1 mediated autoinflammatory diseases: cryopyrin-associated periodic syndromes, tumour necrosis factor receptor-associated periodic syndrome, mevalonate kinase deficiency, and deficiency of the interleukin-1 receptor antagonist. Ann Rheum Dis. 2022;81(7):907–21.

    Article  CAS  PubMed  Google Scholar 

  70. Neven B, Marvillet I, Terrada C, Ferster A, Boddaert N, Couloignier V, et al. Long-term efficacy of the interleukin-1 receptor antagonist anakinra in ten patients with neonatal-onset multisystem inflammatory disease/chronic infantile neurologic, cutaneous, articular syndrome. Arthritis Rheum. 2010;62(1):258–67.

    Article  CAS  PubMed  Google Scholar 

  71. Theodoropoulou K, Spel L, Zaffalon L, Delacretaz M, Hofer M, Martinon F. NLRP3 leucine-rich repeats control induced and spontaneous inflammasome activation in cryopyrin-associated periodic syndrome. J Allergy Clin Immunol. 2023;151(1):222-32 e9.

    Article  CAS  PubMed  Google Scholar 

  72. Ma JH, Lee E, Yoon SH, Min H, Oh JH, Hwang I, et al. Therapeutic effect of NLRP3 inhibition on hearing loss induced by systemic inflammation in a CAPS-associated mouse model. EBioMedicine. 2022;82: 104184.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Chae JJ, Cho YH, Lee GS, Cheng J, Liu PP, Feigenbaum L, et al. Gain-of-function Pyrin mutations induce NLRP3 protein-independent interleukin-1beta activation and severe autoinflammation in mice. Immunity. 2011;34(5):755–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Stoler I, Freytag J, Orak B, Unterwalder N, Henning S, Heim K, et al. Gene-dose effect of MEFV gain-of-function mutations determines ex vivo neutrophil activation in familial Mediterranean fever. Front Immunol. 2020;11:716.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Jorch SK, McNally A, Berger P, Wolf J, Kaiser K, ChetruscaCovash A, et al. Complex regulation of alarmins S100A8/A9 and secretion via gasdermin D pores exacerbates autoinflammation in familial Mediterranean fever. J Allergy Clin Immunol. 2023. https://doi.org/10.1016/j.jaci.2023.01.037.

    Article  PubMed  Google Scholar 

  76. Omenetti A, Carta S, Delfino L, Martini A, Gattorno M, Rubartelli A. Increased NLRP3-dependent interleukin 1beta secretion in patients with familial Mediterranean fever: correlation with MEFV genotype. Ann Rheum Dis. 2014;73(2):462–9.

    Article  CAS  PubMed  Google Scholar 

  77. Mortensen SB, Hansen AE, Mogensen TH, Jakobsen MA, Beck HC, Harvald EB, et al. Pyrin inflammasome activation abrogates interleukin-1 receptor antagonist, suggesting a new mechanism underlying familial Mediterranean fever pathogenesis. Arthritis Rheumatol. 2021;73(11):2116–26.

    Article  CAS  PubMed  Google Scholar 

  78. Kovacs GG. Concepts and classification of neurodegenerative diseases. Handb Clin Neurol. 2017;145:301–7.

    Article  PubMed  Google Scholar 

  79. Panicker N, Kam TI, Wang H, Neifert S, Chou SC, Kumar M, et al. Neuronal NLRP3 is a parkin substrate that drives neurodegeneration in Parkinson’s disease. Neuron. 2022;110(15):2422-37 e9.

    Article  CAS  PubMed  Google Scholar 

  80. Jung ES, Suh K, Han J, Kim H, Kang HS, Choi WS, et al. Amyloid-beta activates NLRP3 inflammasomes by affecting microglial immunometabolism through the Syk-AMPK pathway. Aging Cell. 2022;21(5): e13623.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Pradhan LK, Sahoo PK, Chauhan S, Das SK. Recent advances towards diagnosis and therapeutic fingerprinting for Alzheimer’s disease. J Mol Neurosci. 2022;72(6):1143–65.

    Article  CAS  PubMed  Google Scholar 

  82. Gustavsson A, Norton N, Fast T, Frolich L, Georges J, Holzapfel D, et al. Global estimates on the number of persons across the Alzheimer’s disease continuum. Alzheimers Dement. 2022. https://doi.org/10.1002/alz.12694.

    Article  PubMed  PubMed Central  Google Scholar 

  83. Achudhan D, Liu SC, Lin YY, Huang CC, Tsai CH, Ko CY, et al. Antcin K inhibits TNF-alpha, IL-1beta and IL-8 expression in synovial fibroblasts and ameliorates cartilage degradation: implications for the treatment of rheumatoid arthritis. Front Immunol. 2021;12: 790925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Goldmann T, Tay TL, Prinz M. Love and death: microglia, NLRP3 and the Alzheimer’s brain. Cell Res. 2013;23(5):595–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Zhang Y, Dong Z, Song W. NLRP3 inflammasome as a novel therapeutic target for Alzheimer’s disease. Signal Transduct Target Ther. 2020;5(1):37.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Ising C, Venegas C, Zhang S, Scheiblich H, Schmidt SV, Vieira-Saecker A, et al. NLRP3 inflammasome activation drives tau pathology. Nature. 2019;575(7784):669–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Fang Y, Yao L, Li C, Wang J, Wang J, Chen S, et al. The blockage of the Nogo/NgR signal pathway in microglia alleviates the formation of Abeta plaques and tau phosphorylation in APP/PS1 transgenic mice. J Neuroinflamm. 2016;13(1):56.

    Article  Google Scholar 

  88. Villegas-Llerena C, Phillips A, Garcia-Reitboeck P, Hardy J, Pocock JM. Microglial genes regulating neuroinflammation in the progression of Alzheimer’s disease. Curr Opin Neurobiol. 2016;36:74–81.

    Article  CAS  PubMed  Google Scholar 

  89. Domellof ME, Elgh E, Forsgren L. The relation between cognition and motor dysfunction in drug-naive newly diagnosed patients with Parkinson’s disease. Mov Disord. 2011;26(12):2183–9.

    Article  PubMed  Google Scholar 

  90. Lee AJ, Kim C, Park S, Joo J, Choi B, Yang D, et al. Characterization of altered molecular mechanisms in Parkinson’s disease through cell type-resolved multiomics analyses. Sci Adv. 2023;9(15):eaob2467.

    Article  Google Scholar 

  91. Zheng Z, Lv Y, Rong S, Sun T, Chen L. Physical frailty, genetic predisposition, and incident Parkinson disease. JAMA Neurol. 2023. https://doi.org/10.1001/jamaneurol.2023.0183.

    Article  PubMed  Google Scholar 

  92. Bellomo G, Paciotti S, Concha-Marambio L, Rizzo D, Wojdaƚa AL, Chiasserini D, et al. Cerebrospinal fluid lipoproteins inhibit alpha-synuclein aggregation by interacting with oligomeric species in seed amplification assays. Mol Neurodegener. 2023;18(1):20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Chen X, Feng Y, Quinn RJ, Pountney DL, Richardson DR, Mellick GD, et al. Potassium channels in Parkinson’s disease: potential roles in its pathogenesis and innovative molecular targets for treatment. Pharmacol Rev. 2023. https://doi.org/10.1124/pharmrev.122.000743.

    Article  PubMed  Google Scholar 

  94. Liu Y, Liu T, Zhou Y, Li W, Wang M, Song N, et al. Impeding the combination of astrocytic ASCT2 and NLRP3 by talniflumate alleviates neuroinflammation in experimental models of Parkinson’s disease. Acta Pharm Sin B. 2023;13(2):662–77.

    Article  PubMed  Google Scholar 

  95. Fan Z, Pan YT, Zhang ZY, Yang H, Yu SY, Zheng Y, et al. Systemic activation of NLRP3 inflammasome and plasma alpha-synuclein levels are correlated with motor severity and progression in Parkinson’s disease. J Neuroinflamm. 2020;17(1):11.

    Article  CAS  Google Scholar 

  96. Li Y, Xia Y, Yin S, Wan F, Hu J, Kou L, et al. Targeting microglial alpha-synuclein/TLRs/NF-kappaB/NLRP3 inflammasome axis in Parkinson’s disease. Front Immunol. 2021;12: 719807.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Xie W, Chung KK. Alpha-synuclein impairs normal dynamics of mitochondria in cell and animal models of Parkinson’s disease. J Neurochem. 2012;122(2):404–14.

    Article  CAS  PubMed  Google Scholar 

  98. Gordon R, Albornoz EA, Christie DC, Langley MR, Kumar V, Mantovani S, et al. Inflammasome inhibition prevents alpha-synuclein pathology and dopaminergic neurodegeneration in mice. Sci Transl Med. 2018. https://doi.org/10.1126/scitranslmed.aah4066.

    Article  PubMed  PubMed Central  Google Scholar 

  99. Lee E, Hwang I, Park S, Hong S, Hwang B, Cho Y, et al. MPTP-driven NLRP3 inflammasome activation in microglia plays a central role in dopaminergic neurodegeneration. Cell Death Differ. 2019;26(2):213–28.

    Article  CAS  PubMed  Google Scholar 

  100. Cheng J, Liao Y, Dong Y, Hu H, Yang N, Kong X, et al. Microglial autophagy defect causes Parkinson disease-like symptoms by accelerating inflammasome activation in mice. Autophagy. 2020;16(12):2193–205.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Qin Y, Qiu J, Wang P, Liu J, Zhao Y, Jiang F, et al. Impaired autophagy in microglia aggravates dopaminergic neurodegeneration by regulating NLRP3 inflammasome activation in experimental models of Parkinson’s disease. Brain Behav Immun. 2021;91:324–38.

    Article  CAS  PubMed  Google Scholar 

  102. Ma X, Hao J, Wu J, Li Y, Cai X, Zheng Y. Prussian blue nanozyme as a pyroptosis inhibitor alleviates neurodegeneration. Adv Mater. 2022;34(15): e2106723.

    Article  PubMed  Google Scholar 

  103. Pike AF, Longhena F, Faustini G, van Eik JM, Gombert I, Herrebout MAC, et al. Dopamine signaling modulates microglial NLRP3 inflammasome activation: implications for Parkinson’s disease. J Neuroinflamm. 2022;19(1):50.

    Article  CAS  Google Scholar 

  104. Ahmed S, Kwatra M, Ranjan Panda S, Murty USN, Naidu VGM. Andrographolide suppresses NLRP3 inflammasome activation in microglia through induction of parkin-mediated mitophagy in in-vitro and in-vivo models of Parkinson disease. Brain Behav Immun. 2021;91:142–58.

    Article  CAS  PubMed  Google Scholar 

  105. Keller CG, Shin Y, Monteys AM, Renaud N, Beibel M, Teider N, et al. An orally available, brain penetrant, small molecule lowers huntingtin levels by enhancing pseudoexon inclusion. Nat Commun. 2022;13(1):1150.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Barake F, Soza A, Gonzalez A. Galectins in the brain: advances in neuroinflammation, neuroprotection and therapeutic opportunities. Curr Opin Neurol. 2020;33(3):381–90.

    Article  CAS  PubMed  Google Scholar 

  107. Ren Z, Liang W, Sheng J, Xun C, Xu T, Cao R, et al. Gal-3 is a potential biomarker for spinal cord injury and Gal-3 deficiency attenuates neuroinflammation through ROS/TXNIP/NLRP3 signaling pathway. 2019. Biosci Rep. https://doi.org/10.1042/BSR20192368.

  108. Siew JJ, Chen HM, Chen HY, Chen HL, Chen CM, Soong BW, et al. Galectin-3 is required for the microglia-mediated brain inflammation in a model of Huntington’s disease. Nat Commun. 2019;10(1):3473.

    Article  PubMed  PubMed Central  Google Scholar 

  109. Paldino E, D’Angelo V, Sancesario G, Fusco FR. Pyroptotic cell death in the R6/2 mouse model of Huntington’s disease: new insight on the inflammasome. Cell Death Discov. 2020;6:69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Chen KP, Hua KF, Tsai FT, Lin TY, Cheng CY, Yang DI, et al. 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(1):56.

    Article  CAS  Google Scholar 

  111. Paldino E, D’Angelo V, Laurenti D, Angeloni C, Sancesario G, Fusco FR. Modulation of inflammasome and pyroptosis by olaparib, a PARP-1 inhibitor, in the R6/2 mouse model of Huntington’s disease. Cells. 2020. https://doi.org/10.3390/cells9102286.

    Article  PubMed  PubMed Central  Google Scholar 

  112. Schunk SJ, Kleber ME, Marz W, Pang S, Zewinger S, Triem S, et al. Genetically determined NLRP3 inflammasome activation associates with systemic inflammation and cardiovascular mortality. Eur Heart J. 2021;42(18):1742–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Orecchioni M, Kobiyama K, Winkels H, Ghosheh Y, McArdle S, Mikulski Z, et al. Olfactory receptor 2 in vascular macrophages drives atherosclerosis by NLRP3-dependent IL-1 production. Science. 2022;375(6577):214–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Bjorkegren JLM, Lusis AJ. Atherosclerosis: recent developments. Cell. 2022;185(10):1630–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Gu P, Hui X, Zheng Q, Gao Y, Jin L, Jiang W, et al. Mitochondrial uncoupling protein 1 antagonizes atherosclerosis by blocking NLRP3 inflammasome-dependent interleukin-1beta production. Sci Adv. 2021;7(50): eabl4024.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Duewell P, Kono H, Rayner KJ, Sirois CM, Vladimer G, Bauernfeind FG, et al. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature. 2010;464(7293):1357–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Wang Y, Ji N, Gong X, Ni S, Xu L, Zhang H. Thioredoxin-1 attenuates atherosclerosis development through inhibiting NLRP3 inflammasome. Endocrine. 2020;70(1):65–70.

    Article  CAS  PubMed  Google Scholar 

  118. Wu X, Zhang H, Qi W, Zhang Y, Li J, Li Z, et al. Nicotine promotes atherosclerosis via ROS-NLRP3-mediated endothelial cell pyroptosis. Cell Death Dis. 2018;9(2):171.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Ziaeian B, Fonarow GC. Epidemiology and aetiology of heart failure. Nat Rev Cardiol. 2016;13(6):368–78.

    Article  PubMed  PubMed Central  Google Scholar 

  120. Wang S, Zhang J, Wang Y, Jiang X, Guo M, Yang Z. NLRP3 inflammasome as a novel therapeutic target for heart failure. Anatol J Cardiol. 2022;26(1):15–22.

    Article  CAS  PubMed  Google Scholar 

  121. Zeng C, Duan F, Hu J, Luo B, Huang B, Lou X, et al. NLRP3 inflammasome-mediated pyroptosis contributes to the pathogenesis of non-ischemic dilated cardiomyopathy. Redox Biol. 2020;34: 101523.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Byrne NJ, Matsumura N, Maayah ZH, Ferdaoussi M, Takahara S, Darwesh AM, et al. Empagliflozin blunts worsening cardiac dysfunction associated with reduced NLRP3 (nucleotide-binding domain-like receptor protein 3) inflammasome activation in heart failure. Circ Heart Fail. 2020;13(1): e006277.

    Article  CAS  PubMed  Google Scholar 

  123. Karim MA, Singal AG, Kum HC, Lee YT, Park S, Rich NE, et al. Clinical characteristics and outcomes of nonalcoholic fatty liver disease-associated hepatocellular carcinoma in the United States. Clin Gastroenterol Hepatol. 2022. https://doi.org/10.1016/j.cgh.2022.03.010.

    Article  PubMed  Google Scholar 

  124. Sundaram V, Jalan R, Shah P, Singal AK, Patel AA, Wu T, et al. Acute on chronic liver failure from nonalcoholic fatty liver disease: a growing and aging cohort with rising mortality. Hepatology. 2021;73(5):1932–44.

    Article  PubMed  Google Scholar 

  125. Mridha AR, Wree A, Robertson AAB, Yeh MM, Johnson CD, Van Rooyen DM, et al. NLRP3 inflammasome blockade reduces liver inflammation and fibrosis in experimental NASH in mice. J Hepatol. 2017;66(5):1037–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Yang SJ, Lim Y. Resveratrol ameliorates hepatic metaflammation and inhibits NLRP3 inflammasome activation. Metabolism. 2014;63(5):693–701.

    Article  CAS  PubMed  Google Scholar 

  127. Calcagno DM, Chu A, Gaul S, Taghdiri N, Toomu A, Leszczynska A, et al. NOD-like receptor protein 3 activation causes spontaneous inflammation and fibrosis that mimics human NASH. Hepatology. 2022;76(3):727–41.

    Article  CAS  PubMed  Google Scholar 

  128. Lin Z, Xie X, Gu M, Chen Q, Lu G, Jia X, et al. microRNA-144/451 decreases dendritic cell bioactivity via targeting interferon-regulatory factor 5 to limit DSS-induced colitis. Front Immunol. 2022;13: 928593.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Hsieh YT, Chou YC, Kuo PY, Tsai HW, Yen YT, Shiau AL, et al. Down-regulated miR-146a expression with increased neutrophil extracellular traps and apoptosis formation in autoimmune-mediated diffuse alveolar hemorrhage. J Biomed Sci. 2022;29(1):62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Cheung O, Puri P, Eicken C, Contos MJ, Mirshahi F, Maher JW, et al. Nonalcoholic steatohepatitis is associated with altered hepatic MicroRNA expression. Hepatology. 2008;48(6):1810–20.

    Article  CAS  PubMed  Google Scholar 

  131. Calvente CJ, Tameda M, Johnson CD, Del Pilar H, Lin YC, Adronikou N, et al. Neutrophils contribute to spontaneous resolution of liver inflammation and fibrosis via microRNA-223. J Clin Invest. 2019;129(10):4091–109.

    Article  PubMed  PubMed Central  Google Scholar 

  132. Jimenez Calvente C, Del Pilar H, Tameda M, Johnson CD, Feldstein AE. MicroRNA 223 3p negatively regulates the NLRP3 inflammasome in acute and chronic liver injury. Mol Ther. 2020;28(2):653–63.

    Article  CAS  PubMed  Google Scholar 

  133. Gines P, Krag A, Abraldes JG, Sola E, Fabrellas N, Kamath PS. Liver cirrhosis. Lancet. 2021;398(10308):1359–76.

    Article  PubMed  Google Scholar 

  134. Ma X, Zheng X, Pan L, Zhang X. NLRP3 inflammasome activation in liver cirrhotic patients. Biochem Biophys Res Commun. 2018;505(1):40–4.

    Article  CAS  PubMed  Google Scholar 

  135. Chen P, Zhou YK, Han CS, Chen LJ, Wang YM, Zhuang ZM, et al. Stem cells from human exfoliated deciduous teeth alleviate liver cirrhosis via inhibition of gasdermin D-executed hepatocyte pyroptosis. Front Immunol. 2022;13: 860225.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Barnett R. Liver cirrhosis. Lancet. 2018;392(10144):275.

    Article  PubMed  Google Scholar 

  137. Wang S, Li K, Pickholz E, Dobie R, Matchett KP, Henderson NC, et al. An autocrine signaling circuit in hepatic stellate cells underlies advanced fibrosis in nonalcoholic steatohepatitis. Sci Transl Med. 2023;15(677):eadd3949.

    Article  CAS  PubMed  Google Scholar 

  138. Yaping Z, Ying W, Luqin D, Ning T, Xuemei A, Xixian Y. Mechanism of interleukin-1beta-induced proliferation in rat hepatic stellate cells from different levels of signal transduction. APMIS. 2014;122(5):392–8.

    Article  PubMed  Google Scholar 

  139. Zhu Z, Hu R, Li J, Xing X, Chen J, Zhou Q, et al. Alpinetin exerts anti-inflammatory, anti-oxidative and anti-angiogenic effects through activating the Nrf2 pathway and inhibiting NLRP3 pathway in carbon tetrachloride-induced liver fibrosis. Int Immunopharmacol. 2021;96: 107660.

    Article  CAS  PubMed  Google Scholar 

  140. Kim HY, Choi YJ, Kim SK, Kim H, Jun DW, Yoon K, et al. Auranofin prevents liver fibrosis by system Xc-mediated inhibition of NLRP3 inflammasome. Commun Biol. 2021;4(1):824.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Gaul S, Leszczynska A, Alegre F, Kaufmann B, Johnson CD, Adams LA, et al. Hepatocyte pyroptosis and release of inflammasome particles induce stellate cell activation and liver fibrosis. J Hepatol. 2021;74(1):156–67.

    Article  CAS  PubMed  Google Scholar 

  142. Wree A, Eguchi A, McGeough MD, Pena CA, Johnson CD, Canbay A, et al. NLRP3 inflammasome activation results in hepatocyte pyroptosis, liver inflammation, and fibrosis in mice. Hepatology. 2014;59(3):898–910.

    Article  CAS  PubMed  Google Scholar 

  143. Liu B, Zhou Z, Jin Y, Lu J, Feng D, Peng R, et al. Hepatic stellate cell activation and senescence induced by intrahepatic microbiota disturbances drive progression of liver cirrhosis toward hepatocellular carcinoma. J Immunother Cancer. 2022. https://doi.org/10.1136/jitc-2021-003069.

    Article  PubMed  PubMed Central  Google Scholar 

  144. Bauer C, Duewell P, Mayer C, Lehr HA, Fitzgerald KA, Dauer M, et al. Colitis induced in mice with dextran sulfate sodium (DSS) is mediated by the NLRP3 inflammasome. Gut. 2010;59(9):1192–9.

    Article  CAS  PubMed  Google Scholar 

  145. Zaki MH, Boyd KL, Vogel P, Kastan MB, Lamkanfi M, Kanneganti TD. The NLRP3 inflammasome protects against loss of epithelial integrity and mortality during experimental colitis. Immunity. 2010;32(3):379–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Liu L, Dong Y, Ye M, Jin S, Yang J, Joosse ME, et al. The pathogenic role of NLRP3 inflammasome activation in inflammatory bowel diseases of both mice and humans. J Crohns Colitis. 2017;11(6):737–50.

    PubMed  PubMed Central  Google Scholar 

  147. Villani AC, Lemire M, Fortin G, Louis E, Silverberg MS, Collette C, et al. Common variants in the NLRP3 region contribute to Crohn’s disease susceptibility. Nat Genet. 2009;41(1):71–6.

    Article  CAS  PubMed  Google Scholar 

  148. Mao L, Kitani A, Similuk M, Oler AJ, Albenberg L, Kelsen J, et al. Loss-of-function CARD8 mutation causes NLRP3 inflammasome activation and Crohn’s disease. J Clin Invest. 2018;128(5):1793–806.

    Article  PubMed  PubMed Central  Google Scholar 

  149. Wang K, Lv Q, Miao YM, Qiao SM, Dai Y, Wei ZF. Cardamonin, a natural flavone, alleviates inflammatory bowel disease by the inhibition of NLRP3 inflammasome activation via an AhR/Nrf2/NQO1 pathway. Biochem Pharmacol. 2018;155:494–509.

    Article  CAS  PubMed  Google Scholar 

  150. Tian M, Ma P, Zhang Y, Mi Y, Fan D. Ginsenoside Rk3 alleviated DSS-induced ulcerative colitis by protecting colon barrier and inhibiting NLRP3 inflammasome pathway. Int Immunopharmacol. 2020;85: 106645.

    Article  CAS  PubMed  Google Scholar 

  151. Itani S, Watanabe T, Nadatani Y, Sugimura N, Shimada S, Takeda S, et al. NLRP3 inflammasome has a protective effect against oxazolone-induced colitis: a possible role in ulcerative colitis. Sci Rep. 2016;6:39075.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Yao X, Zhang C, Xing Y, Xue G, Zhang Q, Pan F, et al. Remodelling of the gut microbiota by hyperactive NLRP3 induces regulatory T cells to maintain homeostasis. Nat Commun. 2017;8(1):1896.

    Article  PubMed  PubMed Central  Google Scholar 

  153. Gong Z, He Y, Zhou M, Xin H, Pan M, Fiaz M, et al. Ultrasound imaging tracking of mesenchymal stem cells intracellularly labeled with biosynthetic gas vesicles for treatment of rheumatoid arthritis. Theranostics. 2022;12(5):2370–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Yokota K, Sato K, Miyazaki T, Aizaki Y, Tanaka S, Sekikawa M, et al. Characterization and function of tumor necrosis factor and interleukin-6-induced osteoclasts in rheumatoid arthritis. Arthritis Rheumatol. 2021;73(7):1145–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Cheng L, Liang X, Qian L, Luo C, Li D. NLRP3 gene polymorphisms and expression in rheumatoid arthritis. Exp Ther Med. 2021;22(4):1110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Guo C, Fu R, Wang S, Huang Y, Li X, Zhou M, et al. NLRP3 inflammasome activation contributes to the pathogenesis of rheumatoid arthritis. Clin Exp Immunol. 2018;194(2):231–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Dong X, Zheng Z, Lin P, Fu X, Li F, Jiang J, et al. ACPAs promote IL-1beta production in rheumatoid arthritis by activating the NLRP3 inflammasome. Cell Mol Immunol. 2020;17(3):261–71.

    Article  CAS  PubMed  Google Scholar 

  158. Jager E, Murthy S, Schmidt C, Hahn M, Strobel S, Peters A, et al. Calcium-sensing receptor-mediated NLRP3 inflammasome response to calciprotein particles drives inflammation in rheumatoid arthritis. Nat Commun. 2020;11(1):4243.

    Article  PubMed  PubMed Central  Google Scholar 

  159. Renaudin F, Orliaguet L, Castelli F, Fenaille F, Prignon A, Alzaid F, et al. Gout and pseudo-gout-related crystals promote GLUT1-mediated glycolysis that governs NLRP3 and interleukin-1beta activation on macrophages. Ann Rheum Dis. 2020;79(11):1506–14.

    Article  CAS  PubMed  Google Scholar 

  160. Martinon F, Petrilli V, Mayor A, Tardivel A, Tschopp J. Gout-associated uric acid crystals activate the NALP3 inflammasome. Nature. 2006;440(7081):237–41.

    Article  CAS  PubMed  Google Scholar 

  161. Terkeltaub R, Sundy JS, Schumacher HR, Murphy F, Bookbinder S, Biedermann S, et al. The interleukin 1 inhibitor rilonacept in treatment of chronic gouty arthritis: results of a placebo-controlled, monosequence crossover, non-randomised, single-blind pilot study. Ann Rheum Dis. 2009;68(10):1613–7.

    Article  CAS  PubMed  Google Scholar 

  162. Kingsbury SR, Conaghan PG, McDermott MF. The role of the NLRP3 inflammasome in gout. J Inflamm Res. 2011;4:39–49.

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Goldberg EL, Asher JL, Molony RD, Shaw AC, Zeiss CJ, Wang C, et al. Beta-hydroxybutyrate deactivates neutrophil NLRP3 inflammasome to relieve gout flares. Cell Rep. 2017;18(9):2077–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Lachmann HJ, Kone-Paut I, Kuemmerle-Deschner JB, Leslie KS, Hachulla E, Quartier P, et al. Use of canakinumab in the cryopyrin-associated periodic syndrome. N Engl J Med. 2009;360(23):2416–25.

    Article  CAS  PubMed  Google Scholar 

  165. Kuemmerle-Deschner JB, Hachulla E, Cartwright R, Hawkins PN, Tran TA, Bader-Meunier B, et al. Two-year results from an open-label, multicentre, phase III study evaluating the safety and efficacy of canakinumab in patients with cryopyrin-associated periodic syndrome across different severity phenotypes. Ann Rheum Dis. 2011;70(12):2095–102.

    Article  CAS  PubMed  Google Scholar 

  166. van Hout GP, Bosch L, Ellenbroek GH, de Haan JJ, van Solinge WW, Cooper MA, et al. The selective NLRP3-inflammasome inhibitor MCC950 reduces infarct size and preserves cardiac function in a pig model of myocardial infarction. Eur Heart J. 2017;38(11):828–36.

    PubMed  Google Scholar 

  167. Ren H, Kong Y, Liu Z, Zang D, Yang X, Wood K, et al. Selective NLRP3 (pyrin domain-containing protein 3) inflammasome inhibitor reduces brain injury after intracerebral hemorrhage. Stroke. 2018;49(1):184–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Chen L, Huang CF, Li YC, Deng WW, Mao L, Wu L, et al. Blockage of the NLRP3 inflammasome by MCC950 improves anti-tumor immune responses in head and neck squamous cell carcinoma. Cell Mol Life Sci. 2018;75(11):2045–58.

    Article  CAS  PubMed  Google Scholar 

  169. Wu M, Yang Z, Zhang C, Shi Y, Han W, Song S, et al. Inhibition of NLRP3 inflammasome ameliorates podocyte damage by suppressing lipid accumulation in diabetic nephropathy. Metabolism. 2021;118: 154748.

    Article  CAS  PubMed  Google Scholar 

  170. Coll RC, Hill JR, Day CJ, Zamoshnikova A, Boucher D, Massey NL, et al. MCC950 directly targets the NLRP3 ATP-hydrolysis motif for inflammasome inhibition. Nat Chem Biol. 2019;15(6):556–9.

    Article  CAS  PubMed  Google Scholar 

  171. Coll RC, Robertson AA, Chae JJ, Higgins SC, Munoz-Planillo R, Inserra MC, et al. A small-molecule inhibitor of the NLRP3 inflammasome for the treatment of inflammatory diseases. Nat Med. 2015;21(3):248–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Tapia-Abellan A, Angosto-Bazarra D, Martinez-Banaclocha H, de Torre-Minguela C, Ceron-Carrasco JP, Perez-Sanchez H, et al. MCC950 closes the active conformation of NLRP3 to an inactive state. Nat Chem Biol. 2019;15(6):560–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Youm YH, Nguyen KY, Grant RW, Goldberg EL, Bodogai M, Kim D, et al. The ketone metabolite beta-hydroxybutyrate blocks NLRP3 inflammasome-mediated inflammatory disease. Nat Med. 2015;21(3):263–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Shippy DC, Wilhelm C, Viharkumar PA, Raife TJ, Ulland TK. beta-Hydroxybutyrate inhibits inflammasome activation to attenuate Alzheimer’s disease pathology. J Neuroinflamm. 2020;17(1):280.

    Article  CAS  Google Scholar 

  175. Luo S, Yang M, Han Y, Zhao H, Jiang N, Li L, et al. beta-Hydroxybutyrate against Cisplatin-Induced acute kidney injury via inhibiting NLRP3 inflammasome and oxidative stress. Int Immunopharmacol. 2022;111: 109101.

    Article  CAS  PubMed  Google Scholar 

  176. Corcoran SE, Halai R, Cooper MA. Pharmacological inhibition of the nod-like receptor family pyrin domain containing 3 inflammasome with MCC950. Pharmacol Rev. 2021;73(3):968–1000.

    Article  CAS  PubMed  Google Scholar 

  177. Feng X, Zhan F, Luo D, Hu J, Wei G, Hua F, et al. LncRNA 4344 promotes NLRP3-related neuroinflammation and cognitive impairment by targeting miR-138–5p. Brain Behav Immun. 2021;98:283–98.

    Article  CAS  PubMed  Google Scholar 

  178. Hu H, Wang Y, Ding X, He Y, Lu Z, Wu P, et al. Long non-coding RNA XLOC_000647 suppresses progression of pancreatic cancer and decreases epithelial-mesenchymal transition-induced cell invasion by down-regulating NLRP3. Mol Cancer. 2018;17(1):18.

    Article  PubMed  PubMed Central  Google Scholar 

  179. Richards MW, O’Regan L, Mas-Droux C, Blot JM, Cheung J, Hoelder S, et al. An autoinhibitory tyrosine motif in the cell-cycle-regulated Nek7 kinase is released through binding of Nek9. Mol Cell. 2009;36(4):560–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Kim SK, Choe JY, Park KY. Anti-inflammatory effect of artemisinin on uric acid-induced NLRP3 inflammasome activation through blocking interaction between NLRP3 and NEK7. Biochem Biophys Res Commun. 2019;517(2):338–45.

    Article  CAS  PubMed  Google Scholar 

  181. Chen Y, He H, Lin B, Chen Y, Deng X, Jiang W, et al. RRx-001 ameliorates inflammatory diseases by acting as a potent covalent NLRP3 inhibitor. Cell Mol Immunol. 2021;18(6):1425–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Coll RC, Schroder K, Pelegrin P. NLRP3 and pyroptosis blockers for treating inflammatory diseases. Trends Pharmacol Sci. 2022;43(8):653–68.

    Article  CAS  PubMed  Google Scholar 

  183. Sharma BR, Kanneganti TD. NLRP3 inflammasome in cancer and metabolic diseases. Nat Immunol. 2021;22(5):550–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Tengesdal IW, Menon DR, Osborne DG, Neff CP, Powers NE, Gamboni F, et al. Targeting tumor-derived NLRP3 reduces melanoma progression by limiting MDSCs expansion. Proc Natl Acad Sci U S A. 2021. https://doi.org/10.1073/pnas.2000915118.

    Article  PubMed  PubMed Central  Google Scholar 

  185. Dupaul-Chicoine J, Arabzadeh A, Dagenais M, Douglas T, Champagne C, Morizot A, et al. The Nlrp3 inflammasome suppresses colorectal cancer metastatic growth in the liver by promoting natural killer cell tumoricidal activity. Immunity. 2015;43(4):751–63.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (82070422 and 81871607), Youth Science and Technology Rising Star Project of Shaanxi Province (2020KJXX-036), Innovation Capability Strong Foundation Plan of Xi'an City (Medical Research Project, 21YXYJ0037), Key Research and Development Program of Shaanxi (2020ZDLSF04-03), and Major Research Projects of Xi'an Science and Technology Plan [201805104YX12SF38(2)].

Funding

National Natural Science Foundation of China, 82200330, 82070422, 81871607, Youth Science and Technology Rising Star Project of Shaanxi Province, 2020KJXX-036, Innovation Capability Strong Foundation Plan of Xi'an City, 21YXYJ0037.

Author information

Authors and Affiliations

Authors

Contributions

CY, XYY, and GE collected relevant literature and drafted manuscripts. WRL, ML, and TJ reviewed and made significant revisions to the manuscript. XZ, YY, ZW, and ZYQ prepared figures and tables. YY and DAC guided the preparation of this manuscript. All authors have read and approved the final manuscript.

Corresponding authors

Correspondence to Darío Acuña-Castroviejo or Yang Yang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

All authors agree to publish this review.

Competing interests

The authors have declared that no competing interest exists.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, Y., Ye, X., Escames, G. et al. The NLRP3 inflammasome: contributions to inflammation-related diseases. Cell Mol Biol Lett 28, 51 (2023). https://doi.org/10.1186/s11658-023-00462-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s11658-023-00462-9

Keywords