ArticlePDF AvailableLiterature Review

Loop Between NLRP3 Inflammasome and Reactive Oxygen Species

Authors:

Abstract and Figures

Significance: Inflammasomes are cytosolic multi-protein complexes that mediate innate immune pathways. Inflammasomes activate inflammatory caspases and regulate inflammatory cytokines IL-1β and IL-18 as well as inflammatory cell death (pyroptosis). Among known inflammasomes, NLRP3 inflammasome is unique and well-studied owing to the fact that it senses a broad range of stimuli and is implicated in the pathogenesis of both microbial and sterile inflammatory diseases. Recent Advances: Reactive oxygen species (ROS), especially derived from the mitochondria, are one of the critical mediators of NLRP3 inflammasome activation. Furthermore, NLRP3 inflammasome-driven inflammation recruits inflammatory cells, including macrophages and neutrophils, which in turn cause ROS production, suggesting a feedback loop between ROS and NLRP3 inflammasome. Critical issues: The precise mechanism of how ROS affects NLRP3 inflammasome activation still need to be addressed. This review will summarize the current knowledge on the molecular mechanisms underlying the activation of NLRP3 inflammasome with particular emphasis on the intricate balance of feedback loop between ROS and inflammasome activation. Future directions: Understanding that this relationship is loop rather than traditionally understood linear mechanism will enable to fine tune inflammasome activation under varied pathological settings.
Content may be subject to copyright.
Open camera or QR reader and
scan code to access this article
and other resources online.
Loop Between NLRP3 Inflammasome
and Reactive Oxygen Species
Abishai Dominic,
1,2
Nhat-Tu Le,
2
and Masafumi Takahashi
3,i
Abstract
Significance: Inflammasomes are cytosolic multiprotein complexes that mediate innate immune pathways.
Inflammasomes activate inflammatory caspases and regulate inflammatory cytokines interleukin (IL)-1band
IL-18 as well as inflammatory cell death (pyroptosis). Among known inflammasomes, NLRP3 (NLR family
pyrin domain containing 3) inflammasome is unique and well studied owing to the fact that it senses a broad
range of stimuli and is implicated in the pathogenesis of both microbial and sterile inflammatory diseases.
Recent Advances: Reactive oxygen species (ROS), especially derived from the mitochondria, are one of the
critical mediators of NLRP3 inflammasome activation. Furthermore, NLRP3 inflammasome-driven inflamma-
tion recruits inflammatory cells, including macrophages and neutrophils, which in turn cause ROS production,
suggesting a feedback loop between ROS and NLRP3 inflammasome.
Critical Issues: The precise mechanism of how ROS affects NLRP3 inflammasome activation still need to be
addressed. This review will summarize the current knowledge on the molecular mechanisms underlying the
activation of NLRP3 inflammasome with particular emphasis on the intricate balance of feedback loop between
ROS and inflammasome activation.
Future Directions: Understanding that this relationship is loop rather than traditionally understood linear mech-
anism will enable to fine-tune inflammasome activation under varied pathological settings. Antioxid. Redox Signal.
00, 000–000.
Keywords: cytokines, inflammation, interleukins, macrophages, mitochondria, pyroptosis
Introduction
Inflammation is the response of the body’s innate im-
mune system to harmful stimuli derived from patho-
gens, damaged or dead cells, and irritants. The inflammatory
response is tightly regulated to control its duration and mag-
nitude while preserving host tissue structure and function.
Exposure to harmful stimuli activates multiple signaling cas-
cades to induce the expression of inflammatory cytokines
and chemokines by resident tissue cells. This leads to the
1
Department of Molecular and Cellular Medicine, College of Medicine, Texas A&M University, College Station, Texas, USA.
2
Department of Cardiovascular Sciences, Center for Cardiovascular Regeneration, Houston Methodist Research Institute, Houston,
Texas, USA.
3
Division of Inflammation Research, Center for Molecular Medicine, Jichi Medical University, Tochigi, Japan.
i
ORCID ID (https://orcid.org/0000-0003-2716-7532).
ANTIOXIDANTS & REDOX SIGNALING
Volume 00, Number 00, 2022
ªMary Ann Liebert, Inc.
DOI: 10.1089/ars.2020.8257
1
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
recruitment of inflammatory cells, including macrophages
and neutrophils, inducing a cascade of further production
of substantial amounts of inflammatory cytokines, ultimately
leading to the clearance of the pathogen or stimuli. In addi-
tion, this process also leads to the resolution of inflammation
by clearance of necrotic cells and damaged tissue and initi-
ating the process of tissue repair.
In this manner, inflammatory cytokines are key contribu-
tors to mounting an effective inflammatory response. Inter-
leukins (ILs) are the major type of cytokines that regulate
crucial inflammatory responses. Although >40 cytokines are
designated as ILs, IL-1bis considered the prototypical in-
flammatory cytokine that evokes the secretion of other in-
flammatory cytokines and chemokines, and is implicated in
a wide variety of diseases (54); therefore, targeting of IL-1b-
mediated inflammation has received increasing attention.
One such example is the recent CANTOS (Canakinumab
Anti-inflammatory Thrombosis and Outcome Study) trial
revealed that IL-1binhibition with canakinumab could de-
crease systemic inflammation and cardiovascular events in-
dependent of lipid lowering (75).
A great body of evidence indicates that IL-1bsecretion
is predominantly regulated by intracellular multiprotein com-
plexes called inflammasomes, which create molecular plat-
forms for activating caspase-1 (11, 44, 88). Because caspase-1
is reported to be an IL-1bconverting enzyme (ICE) (47), its
activation drives the processing of pro-IL-1binto active
IL-1b, resulting in inflammation and injury to the surround-
ing tissue. Although there are many different types of in-
flammasomes, the NLRP3 (NLR family pyrin domain
containing 3, also known as NALP3) inflammasome is con-
sidered to be a promising target of various diseases because
its activation is triggered by a broad spectrum of stimuli
derived from pathogens, damaged or dead cells, and irri-
tants, which are known as pathogen-associated molecular
patterns (PAMPs) or damage/danger-associated molecular
patterns (DAMPs).
For this reason, NLRP3 inflammasome has been exten-
sively studied, and it has been revealed that multiple mech-
anisms can participate in NLRP3 inflammasome activation
(44, 53, 58). In particular, reactive oxygen species (ROS)
are one of the main mediators of this activation (44, 88).
In contrast, NLRP3 inflammasome-driven inflammation re-
cruits inflammatory cells, including macrophages and neu-
trophils, which in turn cause ROS production (22, 64). IL-1b
can also promote intracellular accumulation of ROS by dis-
turbing antioxidant enzymes, suggesting a positive feedback
loop between ROS and NLRP3 inflammasome. This review
will summarize the current knowledge on the molecular
mechanisms underlying the activation of NLRP3 inflamma-
some with particular emphasis on the role played by ROS.
NLRP3 Inflammasome
Inflammasomes
The term ‘‘inflammasome’’ was first coined by Martinon
et al. (55) to describe the assembly of an intracellular
multiprotein complex as a molecular platform that triggers
activation of inflammatory caspases (cysteine-dependent
aspartate-specific proteases). The assembly of inflamma-
somes occurs in response to PAMPs and DAMPs. Caspase-1
is a representative inflammatory caspase that regulates the
maturation and secretion of potent pro-inflammatory cyto-
kines such as IL-1b, leading to inflammatory responses (47).
Most inflammasomes typically consist of a sensor protein, the
adaptor protein ASC (apoptosis-associated speck-like protein
containing a caspase-recruitment domain), and caspase-1
(44, 88), although the exact constituents depend on the acti-
vating DAMPs.
Each inflammasome is generally named according to the
identity of the sensor protein. To date, many sensor proteins
that assemble the inflammasomes have been reported: these
include members of the NLR (nucleotide-binding oligo-
merization domain-like receptor) family and the PYHIN
(pyrin and HIN domain) family. Well-known sensor proteins
include NLRP3, NLRP1, NLRC4 (NLR family caspase-
recruitment domain [CARD] containing 4), and AIM2 (ab-
sent in melanoma 2). NLRP1 is activated by lethal toxin and
muramyl dipeptide, whereas NLRC4 is activated by NAIP
(NLR family of apoptosis inhibitory protein) that can rec-
ognize flagellin or components if the bacterial type II secre-
tion system. AIM2 is activated by cytosolic double-strand
DNA.
Among these different types of inflammasomes, NLRP3
is unique and activated by a wide variety of stimuli, including
PAMPs and DAMPs, and is implicated in various diseases
associated with microbial and sterile inflammation, includ-
ing infection, autoinflammatory disorders, cardiovascular
and metabolic diseases, neurological disorders, and cancer
(24, 41, 88, 91, 97). Therefore, NLRP3 inflammasome is the
most intensively studied to date and is considered to be a
promising target for treating a wide variety of diseases.
Regulation of NLRP3 inflammasome assembly
and downstream effects
The NLRP3 inflammasome consists of the sensor NLRP3,
the adaptor ASC, and the effector enzyme pro-caspase-1
(Fig. 1) (44, 88). NLRP3 consists of an N-terminal pyrin
domain (PYD), a central nucleotide-binding and oligomeriza-
tion(NACHT,alsoreferredtoasNOD)domain,anda
C-terminal leucine-rich repeat (LRR) domain. ASC consists of
an N-terminal PYD and a C-terminal CARD. Pro-caspase-1
consists of a CARD and catalytic domains (termed p20/p10
subunits). Although it was previously deemed that the LRR
domain of NLRP3 has autoinhibitory function via an intra-
molecular interaction, a recent study reported that the LRR
domain is dispensable for NLRP3 inflammasome activa-
tion (27).
Upon NLRP3 activation in response to DAMPs or PAMPs,
NLRP3 recruits ASC through PYD–PYD interactions, which
in turn recruits pro-caspase-1 through CARD–CARD inter-
actions, leading to autocleavage of pro-caspase-1 and the
release of active caspase-1 p10 and p20 tetramer, thereby
terminating inflammasome activity (8). The NLRP3 inflam-
masome is mainly expressed in innate immune cells, in-
cluding macrophages and neutrophils, but is also expressed in
nonimmune cells, including epithelial cells, endothelial cells,
and fibroblasts (87).
To assemble the NLRP3 inflammasome and secrete the
active form of IL-1b, a ‘‘two-step’’ signal, the priming signal
(Signal 1) and the activation signal (Signal 2), is required
(Fig. 2) (44, 88). Because the protein levels of NLRP3 and
pro-IL-1bare usually low or undetectable in resting cells,
2 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
a priming signal to increase these protein levels is needed.
This priming signal is mediated at a transcriptional level by
nuclear factor-jB (NF-jB)-via Toll-like receptor (TLR)- or
cytokine receptors. The rapid priming of NLRP3 protein ex-
pression is also mediated by post-translational modification
(PTM), such as de-ubiquitination and phosphorylation (58).
Thereafter, the activation signal promotes the assembly of
NLRP3 inflammasome and induces the activation of caspase-
1 (initially known as ICE) (47). Activated caspase-1 pro-
cesses pro-IL-1b(31 kD) into its active form (17 kD), leading
to the secretion of active IL-1band causing inflammatory
responses. Thus, NLRP3 inflammasome-driven IL-1brelease
is regulated by a ‘‘two-step’’ process. Active caspase-1
also processes pro-IL-18 (inactive, 24 kD) into its active form
(18 kD). However, pro-IL-18 protein is expressed more
constitutively in many types of cells (98), suggesting that
priming for pro-IL-18 may not be necessary. NLRP3
inflammasome-driven activation of caspase-1 also induces
pyroptosis, a lytic form of regulated cell death (37, 63, 82).
Pyroptosis is morphologically distinct from apoptosis and
characterized by cell swelling, plasma membrane rupture,
and the subsequent loss of cytosolic contents. In 2015, gas-
dermin D (GSDMD) was discovered to be a substrate of
caspase-1 and identified as the executioner of pyroptosis (30,
43, 83). The gasdermin was named after the exclusive ex-
pression profile in the gastrointestinal tract and the epithe-
lium of the skin (77), and consists of a family of recently
identified pore-forming effector proteins (9). In addition
FIG. 2. Regulation of NLRP3 inflammasome activation. A two-step signal is required for NLRP3 inflammasome-
driven downstream events. The priming signal (Signal 1) to increase the protein levels of pro-IL-1band NLRP3 is mediated
by NF-jB-mediated signaling via a TLR or cytokine receptors. The rapid priming of NLRP3 protein expression is also
mediated by PTM, such as de-ubiquitination and phosphorylation. The activation signal (Signal 2) promotes the assembly of
NLRP3 inflammasome and induces the activation of caspase-1, leading to the processing of pro-IL-1band pro-IL-18 into
the active forms. Active caspase-1 also induces the processing of GSDMD and its cleaved N-terminus (GSDMD-N) forms
pores on the cell membrane, leading to pyroptotic cell death. Active IL-1band IL-18 are released outside of the cells via the
GSDMD-forming pores, and cause inflammation. GSDMD, gasdermin D; IL, interleukin; NF-jB, nuclear factor-jB; PTM,
post-translational modification; TLR, Toll-like receptor.
FIG. 1. Components and assembly of NLRP3 inflam-
masome. NLRP3 consists of a PYD, a NACHT domain,
and a C-terminal LRR domain. ASC consists of an
N-terminal PYD and a C-terminal CARD. Pro-caspase-1
consists of a CARD and catalytic p20/p10 domains. Upon
NLRP3 activation in response to DAMPs or PAMPs,
NLRP3 recruits ASC through PYD–PYD interactions,
which in turn recruits pro-caspase-1 through CARD–CARD
interactions, leading to autocleavage of pro-caspase-1 and
the formation of active caspase-1 p10 and p20 tetramer.
ASC, apoptosis-associated speck-like protein containing a
caspase-recruitment domain; CARD, caspase-recruitment
domain; DAMP, damage/danger-associated molecular pat-
tern; LRR, leucine-rich repeat; NLRP3, NLR family pyrin
domain containing 3; PAMP, pathogen-associated molecu-
lar pattern; PYD, pyrin domain.
NLRP3 INFLAMMASOME AND ROS 3
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
to processing pro-IL-1band pro-IL-18, active caspase-1 can
process GSDMD into an N-terminal domain (GSDMD-N,
31 kD) and a C-terminal domain (GSDMD-C, 22 kD).
This processing separates autoinhibitory GSDMD-C from
the active GSDMD-N, which translocates to the cell mem-
brane, binds to phosphoinositides (or cardiolipin of bacterial
membrane) on the inner leaflet of the plasma membrane, and
oligomerizes to form membrane pores (inner diameter of 10–
20 nm) that disrupt membrane integrity, thereby leading to
pyroptosis. Concomitantly, the leakage of bioactive IL-1b
and IL-18 as well as intracellular DAMPs, such as high
mobility group box-1 (HMGB1) and S100 proteins, occurs
and this cause inflammatory responses. Because IL-1band
IL-18 lack an N-terminal signal peptide for secretion via the
endoplasmic reticulum (ER) to a Golgi exocytosis pathway,
the mechanism for their secretion was previously unclear.
It is now considered that these cytokines are secreted by at
least two mechanisms. One is secretory autophagy (21, 99)
and the other is gasdermin-mediated pore formation. IL-1b
and IL-18 (*5 nm) are small enough to pass through the
GSDMD-forming pores. Interestingly, a recent study re-
ported that the GSDMD-mediated pores are repaired by en-
dosomal sorting complexes required for transport machinery
and suggested that IL-1bcan be released in the absence of
pyroptotic cell death (76). GSDMD-mediated pyroptosis is
also induced by murine caspase-11 (caspase-4/5 in humans)
(43, 83). Caspase-11 directly detects cytosolic lipopoly-
saccharides (LPS) and promotes GSDMD-mediated pyr-
optosis in the absence of cytokine secretion.
This pathway is known as a noncanonical inflammasome.
Furthermore, Wang et al. (95) recently reported that pyr-
optosis can also be driven by caspase-3-mediated processing
of GSDME (also known as deafness autosomal dominant
nonsyndromic sensorineural 5 [DFNA5]). We also more re-
cently found that NLRP3 inflammasome induces pyroptosis
via an ASC/caspase-8/caspase-3 pathway independent of
caspase-1 (5). Because the pore-forming ability is conserved
throughout gasdermin family proteins (9), it is possible that
other members of the gasdermin family may contribute to
the process of pyroptosis. Accordingly, although pyroptosis
was initially defined as caspase-1-dependent necrotic cell
death, it is now redefined as gasdermin-dependent necrotic
cell death (82).
Activation of NLRP3 inflammasome
Considering that NLRP3 activation is induced by a wide
variety of diverse stimuli and that most NLRP3 activators
have not been shown to directly bind NLRP3, it seems likely
that NLRP3 can sense common cellular events triggered
by the activating stimuli. These cellular events include K
+
efflux (decrease in intracellular K
+
concentration), lyso-
somal dysfunction-induced release of cathepsin B, and
production of ROS, especially from the mitochondria
(Fig. 2) (44, 88). Of these, K
+
efflux is considered to be the
common upstream mediator of NLRP3 inflammasome acti-
vation (67).
However, recent studies have suggested that K
+
efflux is
not a prerequisite for NLRP3 inflammasome activation (26).
In general, lysosomal dysfunction-induced release of ca-
thepsin B is responsible for NLRP3 inflammasome activation
in response to crystals and particles, including monosodium
urate, cholesterol crystals, and silica. ROS are another critical
mediator of NLRP3 inflammasome activation, which is dis-
cussed in detail later in this review.
Several other regulators of NLRP3 inflammasome acti-
vation, including Nek7 (NIMA-related kinase 7) (31, 78, 81),
GBP5 (guanylate-binding protein 5) (80), PKR (double-
stranded RNA-dependent protein kinase) (52), macrophage
migration inhibitory factor (49), and MARK4 (microtubule-
affinity regulating kinase 4) (51), have been reported previ-
ously, but some of them are controversial (44). In addition
to K
+
efflux, other ionic fluxes, including Ca
2+
mobilization,
Na
+
influx, and Cl
-
efflux, can also be involved in NLRP3
inflammasome activation (44). Regarding Cl
-
efflux, chloride
intracellular channels trigger Cl
-
efflux as downstream of
the K
+
efflux-mtROS axis, which activates NLRP3 inflam-
masome via an NLRP3-Nek7 interaction (89).
Furthermore, Chen recently reported that PtdIns4P
(phosphatidylinositol-4-phosphate) on dispersed trans-Golgi
network mediates NLRP3 inflammasome activation (13). To
date, many investigations have shown that the priming and
activation of NLRP3 inflammasome is fine-tuned by nega-
tive and/or positive regulators, including endogenous mod-
ulators (e.g., CARD-only proteins and pyrin-only proteins),
PTMs (e.g., ubiquitination, phosphorylation, nitrosylation, and
sumoylation), and microRNAs (39, 42, 44, 58). The activation
mechanism underlying NLRP3 inflammasome has been re-
vealed to be much more complicated than previously thought
because it appears that the regulation of NLRP3 inflamma-
some varies depending on the types of cells and stimuli
(Table 1). For more information on NLRP3 inflammasome
regulation, see recent excellent reviews (44, 53, 58).
ROS and NLRP3 Inflammasome
ROS are defined as oxygen-containing highly reactive
chemical species with unpaired electron, and mainly include
hydrogen peroxide (H
2
O
2
), superoxide anion (O
2
-
), and
hydroxyl radical (OH
) (18, 64). ROS are produced as a
by-product of oxygen metabolism in the mitochondria,
or produced by cellular enzymes, such as nicotinamide
adenine dinucleotide phosphate (NADPH) oxidases (NOXs),
xanthine oxidase (XO), lipoxygenases (LOXs), and cyclo-
oxygenases (COXs) (12, 18).
Excessive production of ROS causes oxidative damage
to proteins, nucleic acids, and lipids in membrane, and is
generally thought to be harmful; however, ROS also play
an important role in pathogen elimination, especially in
neutrophils and macrophages, and pathophysiological
processes, including cell signaling, proliferation, and dif-
ferentiation (12, 18, 19). Numerous studies using antioxi-
dants (e.g., N-acetyl cysteine [NAC]) support the notion
that ROS are critical mediators of NLRP3 inflammasome
activation (3, 4).
Although previous studies have reported that O
2
-
,H
2
O
2
,
and peroxynitrite (ONOO
-
) play a role in NLRP3 in-
flammasome activation in some cell types, little informa-
tion is available on which species of reactive oxygen induce
the activation (2, 3, 33, 103). ROS also contribute to the
priming signal of NLRP3 inflammasome via ROS-dependent
transcriptional factor NF-jB and LPS-mediated NLRP3
deubiquitination (38). In contrast, NLRP3 inflammasome
activation induces inflammatory responses and recruits
4 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
inflammatory cells, including neutrophils and macrophages,
which also produce ROS, leading to further inflammatory
responses.
Furthermore, IL-1bpromotes intracellular accumulation
of ROS by disturbing superoxide dismutase (SOD) and cat-
alase (29, 57). Thus, it is likely that there is some positive
loop regarding ROS-mediated NLRP3 inflammasome acti-
vation (Fig. 3). However, the role and mechanisms of ROS in
NLRP3 inflammasome activation are still a matter of debate
and ongoing investigation. Furthermore, since NLRP3 in-
flammasome is highly expressed in innate immune cells, in-
cluding macrophages and neutrophils, but is also expressed in
nonimmune cells, including endothelial cells, epithelial cells,
and fibroblasts, the mechanisms of ROS-mediated NLRP3
inflammasome activation may differ depending on cell type.
ROS derived from mitochondria
It is widely accepted that mitochondrial ROS (mtROS) is a
major mediator of NLRP3 inflammasome activation (Fig. 4).
FIG. 3. Proposed mechanisms of ROS-mediated NLRP3 inflammasome activation. Proposed mechanisms of ROS-
mediated NLRP3 inflammasome activation include mtROS, ox-mtDNA, and TXNIP. mtROS are released from damaged
mitochondria and induce NLRP3 inflammasome activation. mtDNA are also released from damaged mitochondria and
oxidized by mtROS or cytosolic ROS (derived from NADPH oxidase), and directly binds NLRP3, leading to NLRP3
inflammasome activation. The newly synthesized mtDNA via a TLR/IRF-1/CMPK2 pathway also contributes to NLRP3
inflammasome activation. Mitophagy-mediated clearance of damaged mitochondria inhibits its activation. We hypothesize
that ROS derived from mitochondria or NADPH oxidase may cause the dissociation of TXNIP from Trx, and induce the
binding of TXNIP with NLRP3, leading to NLRP3 inflammasome activation. IRF-1, interferon regulator factor-1; mtROS,
mitochondrial ROS; NADPH, nicotinamide adenine dinucleotide phosphate; ox-mtDNA, oxidized mtDNA; ROS, reactive
oxygen species; Trx, thioredoxin; TXNIP, thioredoxin-interacting protein.
Table 1. Investigations of Inflammasome Modulation in Preclinical Studies
of Various Pathological Disorders
Disorders investigated Pathway or functional consequence targeted References
Type 2 diabetes Islet mass and insulin resistance (46, 47)
Endothelial dysfunction Inflammation and oxidative stress, pyroptosis,
endothelial cell function and senescence
(48)
Traumatic brain injury and depression Neuroinflammation, chronic stress, microglial
activation, and neuronal cell death
(49, 50)
Cancer Inflammatory microenvironment, IL-1b
dysregulation, oxidative stress, and clonal
expansion
(51, 52)
Cardioprotective signaling IL-6-mediated caspase activation (47, 53)
Pathogenic infection and autoimmune disorders Cytokine production and phagocytosis (54)
IL, interleukin.
NLRP3 INFLAMMASOME AND ROS 5
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
Zhou et al. (104) was the first to demonstrate the role of
mitochondria and mtROS in NLRP3 inflammasome activa-
tion by demonstrating that first, the mitochondrial respiratory
chain complex I inhibitor rotenone and the complex III in-
hibitor antimycin considerably increased mtROS produc-
tion and NLRP3 inflammasome activation. Conversely, the
downregulation of voltage-dependent anion channels by
shRNA decreased mtROS production and NLRP3 inflamma-
some activation.
Second, upon inflammasome stimulation, NLRP3 relo-
cates from ER to mitochondria-associated ER membranes,
and colocalizes with ASC. In this regard, a later study re-
vealed the mechanism of NLRP3 relocation mediated by the
mitochondrial outer membrane protein mitochondrial anti-
viral signaling protein (86). Third, inhibition of autophagy/
mitophagy by 3-MA (3-methyladenine) results in the accu-
mulation of damaged mitochondria and mtROS in parallel
with increased IL-1bproduction. Similarly, downregulation
of autophagy-associated proteins, Atg5 and beclin-1, en-
hances NLRP3 activator-induced IL-1bproduction.
Consistently, Nakahira et al. (69) showed that deficiency of
beclin-1 and another autophagy-associated protein LC3B
promotes mtROS production and inflammasome activation
in macrophages. Interestingly, upon NLRP3 inflammasome
activation, mtDNA is released into the cytosol through in-
creased mitochondrial membrane permeability, and cytosolic
mtDNA can activate AIM2 inflammasome and subsequent IL-
1bproduction. AIM2 is a member of the PRRs family that
recognizes dsDNA and assemblesAIM2 inflammasome (101).
Similarly, Shimada et al. (84) reported that macrophages
lacking mtDNA were unable to secrete IL-1bin response to
NLRP3 activators. As the responsible mechanism, they fur-
ther showed that mtROS lead to oxidized mtDNA that
directly binds NLRP3, leading to activation of NLRP3 in-
flammasome. Intriguingly, Zhong et al. (102) recently re-
ported that TLR signaling promotes the synthesis of mtDNA
by interferon regulator factor-1 (IRF-1) and mitochondrial
deoxyribonucleotide kinase UMP-CMPK2 (CMPK2), and
then the newly synthesized mtDNA is oxidized and drives
NLRP3 inflammasome activation.
Another potential mechanism of mtROS-induced NLRP3
inflammasome activation is thioredoxin-interacting pro-
tein (TXNIP; also known as thioredoxin-binding protein-2
[TBP2] or vitamin D3 upregulated protein 1 [VDUP1]).
TXNIP was originally characterized as a negative regulator
of the antioxidant protein thioredoxin-1 (Trx1) (70). Zhou
et al. (103) found that, under resting conditions, TXNIP binds
to Trx1 and suppresses its antioxidant and thiol-reducing
function. Once excessive ROS are produced, TXNIP is dis-
sociated from oxidized Trx1, which in turn directly binds the
LRR and NACHT domain of the NLRP3, leading to NLRP3
inflammasome activation.
They further confirmed that caspase-1 activation and IL-1b
processing in response to monosodium urate crystals, ATP, and
imiquimod (R-837) were inhibited in bone marrow-derived
macrophages (BMDMs) from TXNIP
-/-
mice. Although Trx1
is a cytosolic protein, it is possible that interaction between
mitochondrial Trx2 and TXNIP may play a similar role in
NLRP3 inflammasome activation. Supporting this idea, Zhou
et al. (104) observed that NLRP3 activators induce TXNIP to
redistribute to mitochondria in a ROS-dependent manner.
Thereafter, numerous studies have demonstrated that mtROS
can act as a critical mediator of NLRP3 inflammasome
activation.
Most of these studies showed that mitochondria-targeting
antioxidants (e.g., MitoTEMPO, SS-31, and MitoQ) had in-
hibitory effects on NLRP3 inflammasome activation and
IL-1bproduction (15, 91, 96). We have previously shown
using an angiotensin II (AII)-infused murine abdominal
aortic aneurysm (AAA) model that mtROS is robustly in-
creased in the adventitial macrophages in the early phase
of AAA formation (91). Moreover, MitoTEMPO or cyclos-
porin A (an inhibitor of mitochondrial permeability transi-
tion) significantly blocked AII-induced IL-1bproduction in
macrophages.
Similarly, the pathogenic role of the mtROS/NLRP3 in-
flammasome pathway has been demonstrated in various ex-
perimental disease models.
ROS derived from NADPH oxidase and other enzymes
The NADPH oxidases are another major source of ROS
production. Earlier studies reported that NLRP3 inflamma-
some activation in response to ATP, silica, and asbestos was
inhibited by diphenyleneiodonium (DPI) and other ROS in-
hibitors (14, 20, 33). In particular, the asbestos-induced
processing of pro-IL-1bwas inhibited in p22
phox
-knockdown
THP-1 cells (20). Based on these observations, it had been
thought that NADPH oxidase-induced ROS could be invol-
ved in NLRP3 inflammasome activation. However, subsequ-
ent studies showed that NADPH oxidases negatively impact
its activation.
Chronic granulomatous disease (CGD) is an inherited dis-
order that is caused by mutations in subunits of the NADPH
oxidase, and characterized by defective ROS production
in phagocytes, resulting in recurrent infections and dysre-
gulated inflammatory responses. Using peripheral blood
mononuclear cells (PBMCs) from CGD patients, several
groups have shown that NADPH oxidase-derived ROS
FIG. 4. Loop between NLRP3 inflammasome and ROS.
ROS activate NLRP3 inflammasome via the priming and
activation signals, and produces IL-1b, which induces in-
flammatory responses and recruits neutrophils and macro-
phages. The accumulated neutrophils and macrophages
produce ROS and lead to further inflammatory responses.
Furthermore, IL-1bpromotes intracellular accumulation of
ROS by disturbing antioxidant enzymes. Thus, our hypo-
thesis is that this phenomenon is due to the likely positive
loop regarding ROS-mediated NLRP3 inflammasome acti-
vation and not a linear activation.
6 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
production did not affect caspase-1 activation or IL-1bse-
cretion in response to known NLRP3 activators (60, 92, 93).
Notably, one study reported that IL-1bsecretion in re-
sponse to uric acid crystals in PBMCs from CGD patients was
fourfold higher than that in PBMCs from control subjects
(93). Consistent with these clinical findings, stretch-induced
caspase-1 activation and IL-1bsecretion were not prevented
in alveolar macrophages deficient in Nox2 subunit gp91
phox
(96). In addition, it has been shown that DPI, commonly used
as a pharmacological inhibitor of NADPH oxidase-derived
ROS, can inhibit mtROS at higher doses (10). These findings
indicate that NADPH oxidases are dispensable for NLRP3
inflammasome activation.
However, many studies report that NADPH oxidase acti-
vation upon inflammasome activation (3). Although the
reason for this contradiction regarding the role of NADPH
oxidases is currently unclear, it might be due to species
variations, different types of cells, spatial-temporal locali-
zation of ROS, and the presence of functionally redundant
Nox proteins (29). In either case, NADPH oxidase is involved
in the inflammasome mediation, but whether it is a conse-
quence, or a trigger is still be delineated. Cross talk between
mtROS and NADPH oxidase (e.g., Nox1, Nox2, and Nox4)
has been demonstrated (16, 46).
In particular, Nox4, localized to the mitochondria, can
contribute to the production of mtROS by inactivation of
mitochondria respiratory chain complex I and uncoupling of
endothelial nitric oxide synthase (45, 50). Regarding Nox4,
Moon et al. (66) recently showed that Nox4 promotes NLRP3
inflammasome activation via fatty acid oxidation. ROS are
also generated by other cytosolic enzymes, including XO,
LOXs, and COXs, and these enzyme-derived ROS have been
shown to contribute to the activation of NLRP3 inflamma-
some (1, 36). In conclusion, there is an undisputed link be-
tween the role of mtROS and NLRP3 activation, whereas the
link between ROS produced by other cellular machinery and
NLRP3 inflammasome still requires clarification.
Antioxidant molecules and species of ROS
Endogenous antioxidant molecules have been reported
to play a role in NLRP3 inflammasome activation. An earlier
study reported that SOD1 deficiency markedly increases
ROS production, and subsequently inhibits caspase-1 acti-
vation and IL-1bsecretion in macrophages. Regarding the
mechanism, higher superoxide production specifically in-
hibits caspase-1 activation by glutathionylation of cysteine
residues of caspase-1 (59). Supporting this, an excessive
amount of ROS has been shown to inactivate caspase-3 (90).
Thus, ROS may act as double-edged sword in NLRP3 in-
flammasome activation depending on their amount. Inter-
estingly, a recent study showed that de-glutathionylation of
Nek7, a regulator of NLRP3 inflammasome, by glutathione
transferase mega 1-1promotes NLRP3 activation (34).
Nrf2 (nuclear factor E2-related factor or nuclear factor
[erythroid-derived 2]-like 2) is a transcription factor that
regulates antioxidant gene expression and acts as a ROS
detoxification factor (32). For this reason, it was previously
anticipated that Nrf2 might have inhibitory effects on NLRP3
inflammasome activation. However, a previous study showed
that Nrf2 deficiency reduces cholesterol crystal-induced in-
flammasome activation and attenuates the development of
atherosclerosis (23). In this regard, Nrf2 has been shown to
upregulate CD36, a scavenger receptor for oxidized low-
density lipoprotein (LDL), in macrophages (35), suggesting
that elevated CD36 levels by Nrf2 potentiate oxidized LDL
incorporation into macrophages, causing the formation of
cholesterol crystals and subsequent NLRP3 inflammasome
activation.
Furthermore, Nrf2 deficiency inhibits caspase-1 activation
and IL-1bsecretion as well as ASC speck formation in re-
sponse to activators of NLRP3 and AIM2, but not NLRC4, in
macrophages (100). Indeed, it has been increasingly reported
that many Chinese herbal medicines that activate Nrf2 can
inhibit NLRP3 inflammasome activation and attenuate dis-
ease severity in inflammation-associated disease models
(32). Although these herbal medicines are not specific to
NRF2, Garstkiewicz et al. (25) recently reported interesting
results that Nrf2 is a positive regulator of the NLRP3 in-
flammasome, whereas Nrf2-activating compounds inhibit
NLRP3 inflammasome activation.
In conclusion, although at first NRF2 was speculated to be
a negative regulator of inflammasome owing to its antioxi-
dant properties as discussed earlier, studies show that the
relationship is not as straightforward as predicted. One of
the major reasons to this is that the pleotropic nature of both
NRF2 and NLRP3. It can also be speculated that under
acute conditions ROS levels can be sustained by NRF2, but
under chronic conditions the effect of NRF2 might be di-
minished leading to be disengaged from inflammasome
pathways. Further studies directed on the acute and chronic
conditions will be able to establish the link between NRF2
and NLRP3.
As mentioned earlier, TXNIP has been shown to function
as a ROS-dependent regulator of NLRP3 inflammasome
activation. Several other studies have reported that TXNIP
plays a similar role in NLRP3 inflammasome activation (1,
65). In contrast, however, Masters et al. (56) showed that
TXNIP
-/-
BMDMs did not influence IL-1bsecretion in re-
sponse to NLRP3 activators, compared with wild-type mac-
rophages. Muri et al. (68) very recently reported that Trx1
promotes NLRP3 inflammasome-driven IL-1bprocessing
and secretion independent of TXNIP. Therefore, it is as-
sumed that ROS-mediated NLRP3 inflammasome activation
is regulated not only by TXNIP/NLRP3 interaction but also
by other unknown mechanisms.
Conclusions and Clinical Perspective
The NLRP3 inflammasome is a pivotal component of the
innate immune system and plays a significant role in the path-
ogenesis of a wide variety of diseases that include infection,
autoinflammatory disorders, cardiovascular and metabolic
diseases, neurological disorders, and cancer. In particular,
sustained activation of NLRP3 inflammasome contributes to
the development of chronic inflammatory cardiovascular
diseases, such as atherosclerosis and AAA.
Although the mechanism of the sustained NLRP3 activation
remains largely unclear, a previous study suggested that the
adenosine A
2A
receptor (A
2A
R) signaling amplifies the prim-
ing and activation signals and increases the duration of NLRP3
inflammasome activation. As the mechanism, the amplification
by A
2A
R signaling is mediated through a cAMP/PKA/CREB/
HIF-1apathway although the inflammasome regulation by
NLRP3 INFLAMMASOME AND ROS 7
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
adenosine does not replace either the priming and activation
signal (72).
In addition, it seems likely that positive loop of NLRP3
inflammasome and ROS participates in the process of the
sustained activation. Moreover, several key emerging con-
cepts of precondition signaling and concepts of hormesis can
be used to explain the redox state and the activation and
inactivation of inflammasomes (85, 94). Specifically, the
dose–response for ROS and subsequent inflammasome acti-
vation is not clearly understood. This becomes crucial in
neurodegenerative diseases where it is believes that cellular
threshold for redox signaling is not adaptive (62, 74).
NLRP3 inflammasome has been extensively studied for
the past decade and substantial knowledge has been gained;
however, it is becoming clearer that the mechanisms regu-
lating NLRP3 inflammasome and how these influence the
pathogenesis of diseases are much more complicated.
Regarding the ROS and NLRP3 inflammasome, several un-
solved questions still need to be addressed. In particular, the
most important question of how ROS affects NLRP3 in-
flammasome activation is not yet fully understood. In gen-
eral, the redox state is involved in wide variety of diseases
and pathologies from neurological to metabolic disorders
(Table 1). As discussed in this review, the loop concept be-
tween ROS and NLRP3 becomes crucial under innate im-
mune activation settings.
In many settings, NLRP3 indirectly is activated by ROS
produced in response to danger signals. Understanding and
fine-tuning the inflammatory response under various condi-
tions will be key in regulating the several pathological con-
ditions. Inhibition of ROS can inhibit NLRP3 activation,
whereas chronic activation of ROS can lead to negative reg-
ulation of NLRP3 (73). Several recent studies have developed
small molecule inhibitors that can target inflammasome at
various stages of the pathway, whereas the specific drug se-
lection can be determined only by understanding the intricate
balance between redox state and inflammasome activation
(79). Thus, understanding the fine balance that governs this
feedback loop is essential in treating chronic conditions.
In contrast, the recent CANTOS trial using canakinumab
validated the concept that IL-1b-driven inflammation di-
rectly contributes to the pathogenesis of atherothrombotic
events and is a new therapeutic option for reducing cardio-
vascular diseases in clinical practice (75); therefore, both
researchers and clinicians should pay more attention to
NLRP3 inflammasome. Because canakinumab is a fully
human anti-IL-1 monoclonal antibody that inhibits IL-1band
has no cross-reactivity with IL-1a, clinical trials using ca-
nakinumab are rather focused on autoinflammatory disorders
such as cryopyrin-associated periodic syndromes with
NLRP3 mutation, rheumatoid arthritis, and gout (48).
However, the widespread use of canakinumab is limited
because of its high cost. In addition, chronic inhibition of
IL-1bmay cause adverse effects such as infection and
impairment of immune homeostasis. Thus, the develop-
ment of less expensive and direct NLRP3 inhibitors is
highly desirable. Hence, a deeper understanding of the
mechanism of NLRP3 inflammasome regulation and the
development of specific NLRP3 inhibitors should not only
offer new therapeutic modalities but also break new
ground for studying the role of NLRP3 inflammasome in
various disorders.
Authors’ Contributions
M.T. conceptualized the content of this study. A.D. and
M.T. reviewed the literature and drafted the article. N.-T.L.
discussed and edited the review.
Acknowledgments
The authors thank past and current members of the
Inflammation Research laboratory for their contributions to
the research.
Author Disclosure Statement
No competing financial interests exist.
Funding Information
This study was supported by grants from the Japan Society
for the Promotion of Science ( JSPS) through Grants-in-Aid
for Scientific Research (18K08112), the Private University
Research Branding Project, the Agency for Medical Research
and Development-Core Research for Evolutional Science
and Technology (AMED-CREST: 18gm0610012h0105).
References
1. Abais JM, Xia M, Li G, Chen Y, Conley SM, Gehr TW,
Boini KM, and Li PL. Nod-like receptor protein 3
(NLRP3) inflammasome activation and podocyte injury
via thioredoxin-interacting protein (TXNIP) during hy-
perhomocysteinemia. J Biol Chem 289: 27159–27168,
2014.
2. Abais JM, Xia M, Li G, Gehr TW, Boini KM, and Li PL.
Contribution of endogenously produced reactive oxygen
species to the activation of podocyte NLRP3 inflamma-
somes in hyperhomocysteinemia. Free Radic Biol Med
67: 211–220, 2014.
3. Abais JM, Xia M, Zhang Y, Boini KM, and Li PL. Redox
regulation of NLRP3 inflammasomes: ROS as trigger or
effector? Antioxid Redox Signal 22: 1111–1129, 2015.
4. Abderrazak A, Syrovets T, Couchie D, El Hadri K, Friguet
B, Simmet T, and Rouis M. NLRP3 inflammasome: from
a danger signal sensor to a regulatory node of oxidative
stress and inflammatory diseases. Redox Biol 4: 296–307,
2015.
5. Aizawa E, Karasawa T, Watanabe S, Komada T, Kimura
H, Kamata R, Ito H, Hishida E, Yamada N, Kasahara T,
Mori Y, and Takahashi M. GSDME-Dependent incom-
plete pyroptosis permits selective IL-1alpha release under
caspase-1 inhibition. iScience 23: 101070, 2020.
6. Bai B, Yang Y, Wang Q, Li M, Tian C, Liu Y, Aung
LHH, Li PF, Yu T, and Chu XM. NLRP3 inflammasome
in endothelial dysfunction. Cell Death Dis 11: 776, 2020.
7. Benetti E, Chiazza F, Patel NS, and Collino M. The
NLRP3 inflammasome as a novel player of the intercel-
lular crosstalk in metabolic disorders. Mediators Inflamm
2013: 678627, 2013.
8. Boucher D, Monteleone M, Coll RC, Chen KW, Ross CM,
Teo JL, Gomez GA, Holley CL, Bierschenk D, Stacey KJ,
Yap AS, Bezbradica JS, and Schroder K. Caspase-1 self-
cleavage is an intrinsic mechanism to terminate inflam-
masome activity. J Exp Med 215: 827–840, 2018.
9. Broz P, Pelegrin P, and Shao F. The gasdermins, a protein
family executing cell death and inflammation. Nat Rev
Immunol 20: 143–157, 2020.
8 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
10. Bulua AC, Simon A, Maddipati R, Pelletier M, Park H,
Kim KY, Sack MN, Kastner DL, and Siegel RM.
Mitochondrial reactive oxygen species promote produc-
tion of proinflammatory cytokines and are elevated in
TNFR1-associated periodic syndrome (TRAPS). J Exp
Med 208: 519–533, 2011.
11. Chan AH and Schroder K. Inflammasome signaling and
regulation of interleukin-1 family cytokines. J Exp Med
217: e20190314, 2020.
12. Chang R, Mamun A, Dominic A, and Le NT. SARS-
CoV-2 mediated endothelial dysfunction: the potential
role of chronic oxidative stress. Front Physiol 11: 605908,
2020.
13. Chen JQ and Chen ZJJ. PtdIns4P on dispersed trans-Golgi
network mediates NLRP3 inflammasome activation.
Nature 564: 71–76, 2018.
14. Cruz CM, Rinna A, Forman HJ, Ventura AL, Persechini
PM, and Ojcius DM. ATP activates a reactive oxygen
species-dependent oxidative stress response and secretion
of proinflammatory cytokines in macrophages. J Biol
Chem 282: 2871–2879, 2007.
15. Dashdorj A, Jyothi KR, Lim S, Jo A, Nguyen MN, Ha J,
Yoon KS, Kim HJ, Park JH, Murphy MP, and Kim SS.
Mitochondria-targeted antioxidant MitoQ ameliorates
experimental mouse colitis by suppressing NLRP3
inflammasome-mediated inflammatory cytokines. BMC
Med 11: 178, 2013.
16. Dikalov S. Cross talk between mitochondria and NADPH
oxidases. Free Radic Biol Med 51: 1289–1301, 2011.
17. This reference has been deleted.
18. Dominic A, Banerjee P, Hamilton DJ, Le N-T, and Abe
J-I. Time-dependent replicative senescence vs. disturbed
flow-induced pre-mature aging in atherosclerosis. Redox
Biol 37: 101614, 2020.
19. Dominic A, Hamilton D, and Abe JI. Mitochondria and
chronic effects of cancer therapeutics: the clinical impli-
cations. J Thromb Thrombolysis 51: 884–889, 2021.
20. Dostert C, Petrilli V, Van Bruggen R, Steele C, Mossman
BT, and Tschopp J. Innate immune activation through
Nalp3 inflammasome sensing of asbestos and silica.
Science 320: 674–677, 2008.
21. Dupont N, Jiang S, Pilli M, Ornatowski W, Bhattacharya
D, and Deretic V. Autophagy-based unconventional se-
cretory pathway for extracellular delivery of IL-1beta.
EMBO J 30: 4701–4711, 2011.
22. Dupre-Crochet S, Erard M, and Nubetae O. ROS pro-
duction in phagocytes: why, when, and where? J Leukoc
Biol 94: 657–670, 2013.
23. Freigang S, Ampenberger F, Spohn G, Heer S, Sham-
shiev AT, Kisielow J, Hersberger M, Yamamoto M,
Bachmann MF, and Kopf M. Nrf2 is essential for
cholesterol crystal-induced inflammasome activation
and exacerbation of atherosclerosis. Eur J Immunol 41:
2040–2051, 2011.
24. Fusco R, Siracusa R, Genovese T, Cuzzocrea S, and Di
Paola R. Focus on the role of NLRP3 inflammasome in
diseases. Int J Mol Sci 21: 4223, 2020.
25. Garstkiewicz M, Strittmatter GE, Grossi S, Sand J, Fenini
G, Werner S, French LE, and Beer HD. Opposing effects
of Nrf2 and Nrf2-activating compounds on the NLRP3
inflammasome independent of Nrf2-mediated gene ex-
pression. Eur J Immunol 47: 806–817, 2017.
26. Gross CJ, Mishra R, Schneider KS, Medard G,
Wettmarshausen J, Dittlein DC, Shi HX, Gorka O, Koenig
PA, Fromm S, Magnani G, Cikovic T, Hartjes L, Smollich
J, Robertson AAB, Cooper MA, Schmidt-Supprian M,
Schuster M, Schroder K, Broz P, Traidl-Hoffmann C,
Beutler B, Kuster B, Ruland J, Schneider S, Perocchi F,
and Gross O. K+Efflux-independent NLRP3 inflamma-
some activation by small molecules targeting mitochon-
dria. Immunity 45: 761–773, 2016.
27. Hafner-Bratkovic I, Susjan P, Lainscek D, Tapia-Abellan
A, Cerovic K, Kadunc L, Angosto-Bazarra D, Pelegrin P,
and Jerala R. NLRP3 lacking the leucine-rich repeat do-
main can be fully activated via the canonical inflamma-
some pathway. Nat Commun 9: 5182, 2018.
28. This reference has been deleted.
29. Harijith A, Ebenezer DL, and Natarajan V. Reactive
oxygen species at the crossroads of inflammasome and
inflammation. Front Physiol 5: 352, 2014.
30. He WT, Wan HQ, Hu LC, Chen PD, Wang X, Huang Z,
Yang ZH, Zhong CQ, and Han JH. Gasdermin D is an
executor of pyroptosis and required for interleukin-1 beta
secretion. Cell Res 25: 1285–1298, 2015.
31. He Y, Zeng MY, Yang DH, Metro B, and Nunez G.
NEK7 is an essential mediator of NLRP3 activation
downstream of potassium efflux. Nature 530: 354–357,
2016.
32. Hennig P, Garstkiewicz M, Grossi S, Di Filippo M,
French LE, and Beer HD. The crosstalk between Nrf2 and
inflammasomes. Int J Mol Sci 19: 562, 2018.
33. Hewinson J, Moore SF, Glover C, Watts AG, and
MacKenzie AB. A key role for redox signaling in rapid
P2X7 receptor-induced IL-1 beta processing in human
monocytes. J Immunol 180: 8410–8420, 2008.
34. Hughes MM, Hooftman A, Angiari S, Tummala P,
Zaslona Z, Runtsch MC, McGettrick AF, Sutton CE,
Diskin C, Rooke M, Takahashi S, Sundararaj S, Casarotto
MG, Dahlstrom JE, Palsson-McDermott EM, Corr SC,
Mills KHG, Preston RJS, Neamati N, Xie Y, Baell JB,
Board PG, and O’Neill LAJ. Glutathione transferase
omega-1 regulates NLRP3 inflammasome activation
through NEK7 deglutathionylation. Cell Rep 29: 151–
161.e5, 2019.
35. Ishii T, Itoh K, Ruiz E, Leake DS, Unoki H, Yamamoto
M, and Mann GE. Role of Nrf2 in the regulation of CD36
and stress protein expression in murine macrophages:
activation by oxidatively modified LDL and 4-
hydroxynonenal. Circ Res 94: 609–616, 2004.
36. Ives A, Nomura J, Martinon F, Roger T, LeRoy D,
Miner JN, Simon G, Busso N, and So A. Xanthine oxi-
doreductase regulates macrophage IL1beta secretion upon
NLRP3 inflammasome activation. Nat Commun 6: 6555,
2015.
37. Jorgensen I and Miao EA. Pyroptotic cell death defends
against intracellular pathogens. Immunol Rev 265: 130–
142, 2015.
38. Juliana C, Fernandes-Alnemri T, Kang S, Farias A, Qin F,
and Alnemri ES. Non-transcriptional priming and deubi-
quitination regulate NLRP3 inflammasome activation.
J Biol Chem 287: 36617–36622, 2012.
39. Karasawa T, Kawashima A, Usui F, Kimura H, Shirasuna
K, Inoue Y, Komada T, Kobayashi M, Mizushina Y,
Sagara J, and Takahashi M. Oligomerized CARD16 pro-
motes caspase-1 assembly and IL-1beta processing. FEBS
Open Bio 5: 348–356, 2015.
40. Kaufmann FN, Costa AP, Ghisleni G, Diaz AP, Rodrigues
ALS, Peluffo H, and Kaster MP. NLRP3 inflammasome-
NLRP3 INFLAMMASOME AND ROS 9
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
driven pathways in depression: clinical and preclinical
findings. Brain Behav Immun 64: 367–383, 2017.
41. Kawaguchi M, Takahashi M, Hata T, Kashima Y, Usui F,
Morimoto H, Izawa A, Takahashi Y, Masumoto J,
Koyama J, Hongo M, Noda T, Nakayama J, Sagara J,
Taniguchi S, and Ikeda U. Inflammasome activation of
cardiac fibroblasts is essential for myocardial ischemia/
reperfusion injury. Circulation 123: 594–604, 2011.
42. Kawashima A, Karasawa T, Tago K, Kimura H, Kamata
R, Usui-Kawanishi F, Watanabe S, Ohta S, Funakoshi-
Tago M, Yanagisawa K, Kasahara T, Suzuki K, and
Takahashi M. ARIH2 ubiquitinates NLRP3 and negatively
regulates NLRP3 inflammasome activation in macro-
phages. J Immunol 199: 3614–3622, 2017.
43. Kayagaki N, Stowe IB, Lee BL, O’Rourke K, Anderson
K, Warming S, Cuellar T, Haley B, Roose-Girma M,
Phung QT, Liu PS, Lill JR, Li H, Wu JS, Kummerfeld S,
Zhang J, Lee WP, Snipas SJ, Salvesen GS, Morris LX,
Fitzgerald L, Zhang YF, Bertram EM, Goodnow CC, and
Dixit VM. Caspase-11 cleaves gasdermin D for non-
canonical inflammasome signalling. Nature 526: 666–671,
2015.
44. Kelley N, Jeltema D, Duan YH, and He Y. The NLRP3
inflammasome: an overview of mechanisms of activation
and regulation. Int J Mol Sci 20:3328, 2019.
45. Koziel R, Pircher H, Kratochwil M, Lener B, Hermann M,
Dencher NA, and Jansen-Durr P. Mitochondrial respira-
tory chain complex I is inactivated by NADPH oxidase
Nox4. Biochem J 452: 231–239, 2013.
46. Kroller-Schon S, Steven S, Kossmann S, Scholz A, Daub
S, Oelze M, Xia N, Hausding M, Mikhed Y, Zinssius E,
Mader M, Stamm P, Treiber N, Scharffetter-Kochanek K,
Li H, Schulz E, Wenzel P, Munzel T, and Daiber A.
Molecular mechanisms of the crosstalk between mito-
chondria and NADPH oxidase through reactive oxygen
species-studies in white blood cells and in animal models.
Antioxid Redox Signal 20: 247–266, 2014.
47. Kuida K, Lippke JA, Ku G, Harding MW, Livingston DJ,
Su MS, and Flavell RA. Altered cytokine export and ap-
optosis in mice deficient in interleukin-1 beta converting
enzyme. Science 267: 2000–2003, 1995.
48. Lachmann HJ, Kone-Paut I, Kuemmerle-Deschner JB, Le-
slie KS, Hachulla E, Quartier P, Gitton X, Widmer A, Patel
N, Hawkins PN; Canakinumab in CAPS Study Group. Use
of canakinumab in the cryopyrin-associated periodic syn-
drome. NEnglJMed360: 2416–2425, 2009.
49. Lang T, Lee JPW, Elgass K, Pinar AA, Tate MD, Aitken
EH, Fan H, Creed SJ, Deen NS, Traore DAK, Mueller I,
Stanisic D, Baiwog FS, Skene C, Wilce MCJ, Mansell A,
Morand EF, and Harris J. Macrophage migration inhibi-
tory factor is required for NLRP3 inflammasome activa-
tion. Nat Commun 9: 2223, 2018.
50. Lee DY, Wauquier F, Eid AA, Roman LJ, Ghosh-
Choudhury G, Khazim K, Block K, and Gorin Y. Nox4
NADPH oxidase mediates peroxynitrite-dependent un-
coupling of endothelial nitric-oxide synthase and fibro-
nectin expression in response to angiotensin II: role of
mitochondrial reactive oxygen species. J Biol Chem 288:
28668–28686, 2013.
51. Li X, Thome S, Ma X, Amrute-Nayak M, Finigan A, Kitt
L, Masters L, James JR, Shi Y, Meng G, and Mallat Z.
MARK4 regulates NLRP3 positioning and inflammasome
activation through a microtubule-dependent mechanism.
Nat Commun 8: 15986, 2017.
52. Lu B, Nakamura T, Inouye K, Li J, Tang Y, Lundback P,
Valdes-Ferrer SI, Olofsson PS, Kalb T, Roth J, Zou
Y, Erlandsson-Harris H, Yang H, Ting JP, Wang H,
Andersson U, Antoine DJ, Chavan SS, Hotamisligil GS,
and Tracey KJ. Novel role of PKR in inflammasome ac-
tivation and HMGB1 release. Nature 488: 670–674, 2012.
53. Mangan MSJ, Olhava EJ, Roush WR, Seidel HM, Glick
GD, and Latz E. Targeting the NLRP3 inflammasome in
inflammatory diseases. Nat Rev Drug Discov 17: 588–606,
2018.
54. Mantovani A, Dinarello CA, Molgora M, and Garlanda C.
Interleukin-1 and related cytokines in the regulation of
inflammation and immunity. Immunity 50: 778–795, 2019.
55. Martinon F, Burns K, and Tschopp J. The inflammasome:
a molecular platform triggering activation of inflamma-
tory caspases and processing of proIL-beta. Mol Cell 10:
417–426, 2002.
56. Masters SL, Dunne A, Subramanian SL, Hull RL,
Tannahill GM, Sharp FA, Becker C, Franchi L, Yoshihara
E, Chen Z, Mullooly N, Mielke LA, Harris J, Coll RC,
Mills KH, Mok KH, Newsholme P, Nunez G, Yodoi J,
Kahn SE, Lavelle EC, and O’Neill LA. Activation of the
NLRP3 inflammasome by islet amyloid polypeptide pro-
vides a mechanism for enhanced IL-1beta in type 2 dia-
betes. Nat Immunol 11: 897–904, 2010.
57. Mathy-Hartert M, Hogge L, Sanchez C, Deby-Dupont G,
Crielaard JM, and Henrotin Y. Interleukin-1beta and
interleukin-6 disturb the antioxidant enzyme system in bovine
chondrocytes: a possible explanation for oxidative stress
generation. Osteoarthritis Cartilage 16: 756–763, 2008.
58. McKee CM and Coll RC. NLRP3 inflammasome priming:
a riddle wrapped in a mystery inside an enigma. J Leukoc
Biol 108: 937–952, 2020.
59. Meissner F, Molawi K, and Zychlinsky A. Superoxide
dismutase 1 regulates caspase-1 and endotoxic shock. Nat
Immunol 9: 866–872, 2008.
60. Meissner F, Seger RA, Moshous D, Fischer A,
Reichenbach J, and Zychlinsky A. Inflammasome activa-
tion in NADPH oxidase defective mononuclear phago-
cytes from patients with chronic granulomatous disease.
Blood 116: 1570–1573, 2010.
61. This reference has been deleted.
62. Miquel S, Champ C, Day J, Aarts E, Bahr BA, Bakker M,
Banati D, Calabrese V, Cederholm T, Cryan J, Dye L,
Farrimond JA, Korosi A, Laye S, Maudsley S, Milenkovic
D, Mohajeri MH, Sijben J, Solomon A, Spencer JPE,
Thuret S, Vanden Berghe W, Vauzour D, Vellas B,
Wesnes K, Willatts P, Wittenberg R, and Geurts L. Poor
cognitive ageing: vulnerabilities, mechanisms and the
impact of nutritional interventions. Ageing Res Rev 42:
40–55, 2018.
63. Mishra PK, Adameova A, Hill JA, Baines CP, Kang PM,
Downey JM, Narula J, Takahashi M, Abbate A, Piristine
HC, Kar S, Su S, Higa JK, Kawasaki NK, and Matsui T.
Guidelines for evaluating myocardial cell death. Am J
Physiol Heart Circ Physiol 317: H891–H922, 2019.
64. Mittal M, Siddiqui MR, Tran K, Reddy SP, and Malik AB.
Reactive oxygen species in inflammation and tissue in-
jury. Antioxid Redox Signal 20: 1126–1167, 2014.
65. Mohamed IN, Hafez SS, Fairaq A, Ergul A, Imig JD, and
El-Remessy AB. Thioredoxin-interacting protein is re-
quired for endothelial NLRP3 inflammasome activation
and cell death in a rat model of high-fat diet. Diabetologia
57: 413–423, 2014.
10 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
66. Moon JS, Nakahira K, Chung KP, DeNicola GM,
Koo MJ, Pabon MA, Rooney KT, Yoon JH, Ryter
SW, Stout-Delgado H, and Choi AM. NOX4-dependent
fatty acid oxidation promotes NLRP3 inflammasome
activation in macrophages. Nat Med 22: 1002–1012,
2016.
67. Munoz-Planillo R, Kuffa P, Martinez-Colon G, Smith BL,
Rajendiran TM, and Nunez G. K+efflux is the common
trigger of NLRP3 inflammasome activation by bacterial
toxins and particulate matter. Immunity 38: 1142–1153,
2013.
68. Muri J, Thut H, Feng Q, and Kopf M. Thioredoxin-1
distinctly promotes NF-kappaB target DNA binding
and NLRP3 inflammasome activation independently of
Txnip. Elife 9:e53627, 2020.
69. Nakahira K, Haspel JA, Rathinam VA, Lee SJ, Dolinay T,
Lam HC, Englert JA, Rabinovitch M, Cernadas M, Kim
HP, Fitzgerald KA, Ryter SW, and Choi AM. Autophagy
proteins regulate innate immune responses by inhibiting
the release of mitochondrial DNA mediated by the
NALP3 inflammasome. Nat Immunol 12: 222–230, 2011.
70. Nishiyama A, Matsui M, Iwata S, Hirota K, Masutani H,
Nakamura H, Takagi Y, Sono H, Gon Y, and Yodoi J.
Identification of thioredoxin-binding protein-2/vitamin
D(3) up-regulated protein 1 as a negative regulator of
thioredoxin function and expression. J Biol Chem 274:
21645–21650, 1999.
71. This reference has been deleted.
72. Ouyang XS, Ghani A, Malik A, Wilder T, Colegio OR,
Flavell RA, Cronstein BN, and Mehal WZ. Adenosine is
required for sustained inflammasome activation via the
A(2A) receptor and the HIF-1 alpha pathway. Nat Com-
mun 4: 2909, 2013.
73. Pelegrin P and Surprenant A. Dynamics of macrophage
polarization reveal new mechanism to inhibit IL-1beta
release through pyrophosphates. EMBO J 28: 2114–2127,
2009.
74. Pennisi M, Crupi R, Di Paola R, Ontario ML, Bella R,
Calabrese EJ, Crea R, Cuzzocrea S, and Calabrese V.
Inflammasomes, hormesis, and antioxidants in neuro-
inflammation: role of NRLP3 in Alzheimer disease.
J Neurosci Res 95: 1360–1372, 2017.
75. Ridker PM, Everett BM, Thuren T, MacFadyen JG, Chang
WH, Ballantyne C, Fonseca F, Nicolau J, Koenig W,
Anker SD, Kastelein JJP, Cornel JH, Pais P, Pella D,
Genest J, Cifkova R, Lorenzatti A, Forster T, Kobalava Z,
Vida-Simiti L, Flather M, Shimokawa H, Ogawa H,
Dellborg M, Rossi PRF, Troquay RPT, Libby P, Glynn
RJ; CANTOS Trial Group. Antiinflammatory therapy with
canakinumab for atherosclerotic disease. N Engl J Med
377: 1119–1131, 2017.
76. Ruhl S, Shkarina K, Demarco B, Heilig R, Santos JC, and
Broz P. ESCRT-dependent membrane repair negatively
regulates pyroptosis downstream of GSDMD activation.
Science 362: 956–960, 2018.
77. Saeki N, Kuwahara Y, Sasaki H, Satoh H, and Shiroishi T.
Gasdermin (Gsdm) localizing to mouse Chromosome 11
is predominantly expressed in upper gastrointestinal tract
but significantly suppressed in human gastric cancer cells.
Mamm Genome 11: 718–724, 2000.
78. Schmid-Burgk JL, Chauhan D, Schmidt T, Ebert TS,
Reinhardt J, Endl E, and Hornung V. A genome-wide
CRISPR (Clustered Regularly Interspaced Short Palin-
dromic Repeats) screen identifies NEK7 as an essential
component of NLRP3 inflammasome activation. J Biol
Chem 291: 103–109, 2016.
79. Schwaid AG and Spencer KB. Strategies for targeting the
NLRP3 inflammasome in the clinical and preclinical
space. J Med Chem 64: 101–122, 2021.
80. Shenoy AR, Wellington DA, Kumar P, Kassa H, Booth
CJ, Cresswell P, and MacMicking JD. GBP5 promotes
NLRP3 inflammasome assembly and immunity in mam-
mals. Science 336: 481–485, 2012.
81. Shi HX, Wang Y, Li XH, Zhan XM, Tang M, Fina M,
Su LJ, Pratt D, Bu CH, Hildebrand S, Lyon S, Scott L,
Quan JX, Sun QH, Russell J, Arnett S, Jurek P, Chen D,
Kravchenko VV, Mathison JC, Moresco EMY, Monson
NL, Ulevitch RJ, and Beutler B. NLRP3 activation and
mitosis are mutually exclusive events coordinated by
NEK7, a new inflammasome component. Nat Immunol 17:
250–258, 2016.
82. Shi JJ, Gao WQ, and Shao F. Pyroptosis: gasdermin-
mediated programmed necrotic cell death. Trends Bio-
chem Sci 42: 245–254, 2017.
83. Shi JJ, Zhao Y, Wang K, Shi XY, Wang Y, Huang HW,
Zhuang YH, Cai T, Wang FC, and Shao F. Cleavage of
GSDMD by inflammatory caspases determines pyroptotic
cell death. Nature 526: 660–665, 2015.
84. Shimada K, Crother TR, Karlin J, Dagvadorj J, Chiba N,
Chen S, Ramanujan VK, Wolf AJ, Vergnes L, Ojcius
DM, Rentsendorj A, Vargas M, Guerrero C, Wang Y,
Fitzgerald KA, Underhill DM, Town T, and Arditi M.
Oxidized mitochondrial DNA activates the NLRP3
inflammasome during apoptosis. Immunity 36: 401–414,
2012.
85. Siracusa R, Scuto M, Fusco R, Trovato A, Ontario ML,
Crea R, Di Paola R, Cuzzocrea S, and Calabrese V. Anti-
inflammatory and anti-oxidant activity of Hidrox((R))
in rotenone-induced Parkinson’s disease in mice. Anti-
oxidants (Basel) 9: 824, 2020.
86. Subramanian N, Natarajan K, Clatworthy MR, Wang Z,
and Germain RN. The adaptor MAVS promotes NLRP3
mitochondrial localization and inflammasome activation.
Cell 153: 348–361, 2013.
87. Takahashi M. Cell-specific roles of NLRP3 inflammasome
in myocardial infarction. J Cardiovasc Pharm 74: 188–
193, 2019.
88. Takahashi M. NLRP3 inflammasome as a novel player in
myocardial infarction. Int Heart J 55: 101–105, 2014.
89. Tang T, Lang X, Xu C, Wang X, Gong T, Yang Y, Cui J,
Bai L, Wang J, Jiang W, and Zhou R. CLICs-dependent
chloride efflux is an essential and proximal upstream
event for NLRP3 inflammasome activation. Nat Commun
8: 202, 2017.
90. Ueda S, Nakamura H, Masutani H, Sasada T, Yonehara S,
Takabayashi A, Yamaoka Y, and Yodoi J. Redox regu-
lation of caspase-3(-like) protease activity: regulatory
roles of thioredoxin and cytochrome c. J Immunol 161:
6689–6695, 1998.
91. Usui F, Shirasuna K, Kimura H, Tatsumi K, Kawashima
A, Karasawa T, Yoshimura K, Aoki H, Tsutsui H, Noda T,
Sagara J, Taniguchi S, and Takahashi M. Inflammasome
activation by mitochondrial oxidative stress in macro-
phages leads to the development of angiotensin II-induced
aortic aneurysm. Arterioscler Thromb Vasc Biol 35: 127–
136, 2015.
92. van Bruggen R, Koker MY, Jansen M, van Houdt M, Roos
D, Kuijpers TW, and van den Berg TK. Human NLRP3
NLRP3 INFLAMMASOME AND ROS 11
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
inflammasome activation is Nox1–Nox4 independent.
Blood 115: 5398–5400, 2010.
93. van de Veerdonk FL, Smeekens SP, Joosten LA, Kullberg
BJ, Dinarello CA, van der Meer JW, and Netea MG.
Reactive oxygen species-independent activation of the
IL-1beta inflammasome in cells from patients with
chronic granulomatous disease. Proc Natl Acad Sci U S A
107: 3030–3033, 2010.
94. Vittorio Calabrese CC, Albena T. Dinkova-Kostova,
Edward J. Calabrese, and Mark P. Mattson. Cellular stress
responses, the Hormesis paradigm, and vitagenes: novel
targets for therapeutic intervention in neurodegenerative
disorders. Antioxid Redox Signal 13: 1763–1811, 2010.
95. Wang Y, Gao W, Shi X, Ding J, Liu W, He H, Wang K,
and Shao F. Chemotherapy drugs induce pyroptosis
through caspase-3 cleavage of a gasdermin. Nature 547:
99–103, 2017.
96. Wu J, Yan Z, Schwartz DE, Yu J, Malik AB, and Hu G.
Activation of NLRP3 inflammasome in alveolar macro-
phages contributes to mechanical stretch-induced lung in-
flammation and injury. J Immunol 190: 3590–3599, 2013.
97. Yajima N, Takahashi M, Morimoto H, Shiba Y, Takahashi
Y, Masumoto J, Ise H, Sagara J, Nakayama J, Taniguchi
S, and Ikeda U. Critical role of bone marrow apoptosis-
associated speck-like protein, an inflammasome adaptor
molecule, in neointimal formation after vascular injury in
mice. Circulation 117: 3079–3087, 2008.
98. Yasuda K, Nakanishi K, and Tsutsui H. Interleukin-18 in
health and disease. Int J Mol Sci 20: 649, 2019.
99. Zhang M, Kenny SJ, Ge L, Xu K, and Schekman R.
Translocation of interleukin-1beta into a vesicle interme-
diate in autophagy-mediated secretion. Elife 4: e11205,
2015.
100. Zhao C, Gillette DD, Li X, Zhang Z, and Wen H. Nuclear
factor E2-related factor-2 (Nrf2) is required for NLRP3
and AIM2 inflammasome activation. J Biol Chem 289:
17020–17029, 2014.
101. Zheng D, Liwinski T, and Elinav E. Inflammasome acti-
vation and regulation: toward a better understanding of
complex mechanisms. Cell Discov 6: 36, 2020.
102. Zhong Z, Liang S, Sanchez-Lopez E, He F, Shalapour S,
Lin XJ, Wong J, Ding S, Seki E, Schnabl B, Hevener AL,
Greenberg HB, Kisseleva T, and Karin M. New mito-
chondrial DNA synthesis enables NLRP3 inflammasome
activation. Nature 560: 198–203, 2018.
103. Zhou R, Tardivel A, Thorens B, Choi I, and Tschopp J.
Thioredoxin-interacting protein links oxidative stress to
inflammasome activation. Nat Immunol 11: 136–140,
2010.
104. Zhou R, Yazdi AS, Menu P, and Tschopp J. A role for
mitochondria in NLRP3 inflammasome activation. Nature
469: 221–225, 2011.
105. This reference has been deleted.
Address correspondence to:
Dr. Masafumi Takahashi
Division of Inflammation Research
Center for Molecular Medicine
Jichi Medical University
3311-1 Yakushiji, Shimotsuke
Tochigi 329-0498
Japan
E-mail: masafumi2@jichi.ac.jp
Date of first submission to ARS Central, December 24, 2020;
date of final revised submission, September 2, 2021; date of
acceptance, September 3, 2021.
Abbreviations Used
3-MA ¼3-methyladenine
A
2A
R¼adenosine A
2A
receptor
AAA ¼abdominal aortic aneurysm
AII ¼angiotensin II
AIM2 ¼absent in melanoma 2
ASC ¼apoptosis-associated speck-like protein
containing a caspase-recruitment domain
BMDM ¼bone marrow-derived macrophage
CANTOS ¼Canakinumab Anti-inflammatory
Thrombosis and Outcome Study
CARD ¼caspase-recruitment domain
CGD ¼chronic granulomatous disease
COX ¼cyclooxygenase
CREB ¼cAMP response element binding protein
DAMP ¼damage/danger-associated molecular
pattern
DFNA5 ¼deafness autosomal dominant
nonsyndromic sensorineural 5
DPI ¼diphenyleneiodonium
ER ¼endoplasmic reticulum
GBP5 ¼guanylate-binding protein 5
GSDMD ¼gasdermin D
H
2
O
2
¼hydrogen peroxide
HIF-1a¼hypoxia-inducible factor-1a
HMGB-1 ¼high mobility group box-1
ICE ¼IL-1bconverting enzyme
IL ¼interleukin
IRF-1 ¼interferon regulator factor-1
LDL ¼low-density lipoprotein
LOX ¼lipoxygenase
LPS ¼lipopolysaccharides
LRR ¼leucine-rich repeat
MARK4 ¼microtubule-affinity regulating kinase 4
mtROS ¼mitochondrial ROS
NAC ¼N-acetyl cysteine
NADPH ¼nicotinamide adenine dinucleotide phosphate
NAIP ¼NLR family of apoptosis inhibitory protein
Nek7 ¼NIMA-related kinase 7
NF-jB¼nuclear factor-jB
NLR ¼nucleotide-binding oligomerization
domain-like receptor
NLRC4 ¼NLR family caspase-recruitment
domain containing 4
NLRP3 ¼NLR family pyrin domain containing 3
NOX ¼NADPH oxidase
Nrf2 ¼nuclear factor E2-related factor or nuclear
factor [erythroid-derived 2]-like 2
O
2
-
¼superoxide anion
OH
¼hydroxyl radical
ONOO
-
¼peroxynitrite
PAMP ¼pathogen-associated molecular pattern
PBMC ¼peripheral blood mononuclear cell
PKR ¼double-stranded RNA-dependent protein
kinase
PtdIns4P ¼phosphatidylinositol-4-phosphate
PTM ¼post-translational modification
PYD ¼pyrin domain
12 DOMINIC ET AL.
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
Abbreviations Used (Cont.)
PYHIN ¼pyrin and HIN domain
ROS ¼reactive oxygen species
SOD ¼superoxide dismutase
TBP2 ¼thioredoxin-binding protein-2
TLR ¼Toll-like receptor
Trx ¼thioredoxin
TXNIP ¼thioredoxin-interacting protein
VDUP1 ¼vitamin D3 upregulated protein 1
XO ¼xanthine oxidase
NLRP3 INFLAMMASOME AND ROS 13
Downloaded by Texas A&M University College Station from www.liebertpub.com at 02/09/22. For personal use only.
... The intracellular localization of NLRP3 is necessary to mediate a rapid response to multiple cellular danger signals, such as K + efflux, Ca 2+ mobilization, and reactive oxygen species (ROS) production, which have been proposed to drive NLRP3 inflammasome activation 17 . ROS have been speculated to be central and shared common signaling molecules, with their production triggered by a wide range of stimuli to mediate NLRP3 inflammasome activation by regulating both priming and activation signals [18][19][20][21] . Electron transport in mitochondria and NADPH oxidase (NOX) in the membrane are two major drivers of ROS production, and thioredoxin-interacting protein (TXNIP) and oxidized mitochondrial DNA (mtDNA) have been reported to mediate mitochondrial ROS-triggered NLRP3 inflammasome activation 22,23 . ...
... NADPH oxidase was originally characterized as critical for NLRP3 inflammasome activation when ROS inhibitor treatment was used and in p22 phox -knockdown cells 28 . However, caspase-1 activation and IL-1β release were reported to be normal and even increased in macrophages from NADPH oxidase-impaired patients or mice with NADPH oxidase component deficiency, including low levels of membrane-bound p22 phox and NOX2 (gp91 phox ) and cytosolic NCF2 (p67 phox ) 20,26,29,[49][50][51] . These studies by different groups consistently indicate the dispensable roles of NADPH oxidases in inflammasome activation. ...
Article
Full-text available
The spatiotemporal regulation of inflammasome activation remains unclear. To examine the mechanism underlying the assembly and regulation of the inflammasome response, here we perform an immunoprecipitation-mass spectrometry analysis of apoptosis-associated speck-like protein containing a CARD (ASC) and identify NCF4/1/2 as ASC-binding proteins. Reduced NCF4 expression is associated with colorectal cancer development and decreased five-year survival rate in patients with colorectal cancer. NCF4 cooperates with NCF1 and NCF2 to promote NLRP3 and AIM2 inflammasome activation. Mechanistically, NCF4 phosphorylation and puncta distribution switches from the NADPH complex to the perinuclear region, mediating ASC oligomerization, speck formation and inflammasome activation. NCF4 functions as a sensor of ROS levels, to establish a balance between ROS production and inflammasome activation. NCF4 deficiency causes severe colorectal cancer in mice, increases transit-amplifying and precancerous cells, reduces the frequency and activation of CD8⁺ T and NK cells, and impairs the inflammasome-IL-18-IFN-γ axis during the early phase of colorectal tumorigenesis. Our study implicates NCF4 in determining the spatial positioning of inflammasome assembly and contributing to inflammasome-mediated anti-tumor responses.
... Therefore, we used LPS/AβO model according to previous studies. NLRP3 inflammasome is an intracellular multiprotein complex composed of NLRP3, apoptosis-associated speck-like protein containing a caspase recruitment domain (ASC), and cysteine protease Caspase-1, and contributes to inflammatory responses in various diseases, including cardiovascular diseases (Takahashi 2022). The formation of NLRP3 inflammasome causes auto-activation of Caspase-1 which leads to the processing of pro-IL-1β and pro-IL-18 to their forms, resulting in inflammation (Huang et al. 2021). ...
Article
Full-text available
Neuroinflammation is a key factor in cognitive dysfunction and neurodegenerative diseases such as Alzheimer’s disease (AD), so inhibiting neuroinflammation is considered as a potential treatment for AD. Epigallocatechin-3-gallate (EGCG), a polyhydroxyphenol of green tea, has been found to exhibit anti-oxidative, anti-inflammatory and neuroprotective effects. The aim of this study was to investigate the inhibitory effect of EGCG on inflammation and its mechanism. In this study, BV2 cells were simultaneously exposed to lipopolysaccharides (LPS) and the amyloid-β oligomer (AβO) to induce inflammatory microenvironments. Inflammatory cytokines and NLRP3 inflammasome-related molecules were detected by RT-PCR and Western Blot. The results show that EGCG inhibits LPS/AβO-induced inflammation in BV2 cells through regulating IL-1β, IL-6, and TNF-α. Meanwhile, EGCG reduces the activation of the NOD-, LRR-, and pyrin domain-containing protein 3 (NLRP3) inflammasome and levels of intracellular ROS in BV2 cells treated with LPS/AβO by affecting the mitochondrial membrane potential (MMP). Further research found that EGCG inhibited MMP through regulating thioredoxin-interacting protein (TXNIP) in LPS/AβO-induced neuroinflammation. In conclusion, EGCG may alleviate LPS/AβO-induced microglial neuroinflammation by suppressing the ROS/ TXNIP/ NLRP3 pathway. It may provide a potential mechanism underlying the anti-inflammatory properties of EGCG for alleviating AD.
... However, increased ROS levels lead to the disassociation of TXNIP from Trx1 and facilitates the interaction between TXNIP and NLRP3, activating NLRP3. Additionally, mt-ROS-targeting antioxidants inhibit NLRP3 activation [99][100][101]. All these observations collectively establish a crucial role of mt-ROS in the activation of innate immune effector responses. Figure 4. mt-ROS in NLRP3 inflammasome formation. ...
Article
Full-text available
Reactive oxygen species (ROS) contain at least one oxygen atom and one or more unpaired electrons and include singlet oxygen, superoxide anion radical, hydroxyl radical, hydroperoxyl radical, and free nitrogen radicals. Intracellular ROS can be formed as a consequence of several factors, including ultra-violet (UV) radiation, electron leakage during aerobic respiration, inflammatory responses mediated by macrophages, and other external stimuli or stress. The enhanced production of ROS is termed oxidative stress and this leads to cellular damage, such as protein carbonylation, lipid peroxidation, deoxyribonucleic acid (DNA) damage, and base modifications. This damage may manifest in various pathological states, including ageing, cancer, neurological diseases, and metabolic disorders like diabetes. On the other hand, the optimum levels of ROS have been implicated in the regulation of many important physiological processes. For example, the ROS generated in the mitochondria (mitochondrial ROS or mt-ROS), as a byproduct of the electron transport chain (ETC), participate in a plethora of physiological functions, which include ageing, cell growth, cell proliferation, and immune response and regulation. In this current review, we will focus on the mechanisms by which mt-ROS regulate different pathways of host immune responses in the context of infection by bacteria, protozoan parasites, viruses, and fungi. We will also discuss how these pathogens, in turn, modulate mt-ROS to evade host immunity. We will conclude by briefly giving an overview of the potential therapeutic approaches involving mt-ROS in infectious diseases.
... ROS may in part be the link between the two pathways. ROS act as agonists to activate NLRP3 inflammasome [56,57], whereas Nrf2 responds to endogenous and exogenous stress induced by ROS or electrophiles by inducing antioxidant Nrf2 target genes, such as HO-1 and NAD(P)H: quinone oxidoreductase 1, to eliminate oxidative stress and inflammation [57,58]. Recent studies have shown that natural phytochemicals such as magnolol, gallic acid, and rhein can Fig. 7 Butein ameliorates LPS-induced peritonitis. ...
Article
Full-text available
Background Aberrant inflammatory responses drive the initiation and progression of various diseases, and hyperactivation of NLRP3 inflammasome is a key pathogenetic mechanism. Pharmacological inhibitors of NLRP3 represent a potential therapy for treating these diseases but are not yet clinically available. The natural product butein has excellent anti-inflammatory activity, but its potential mechanisms remain to be investigated. In this study, we aimed to evaluate the ability of butein to block NLRP3 inflammasome activation and the ameliorative effects of butein on NLRP3-driven diseases. Methods Lipopolysaccharide (LPS)-primed bone-marrow-derived macrophages were pretreated with butein and various inflammasome stimuli. Intracellular potassium levels, ASC oligomerization and reactive oxygen species production were also detected to evaluate the regulatory mechanisms of butein. Moreover, mouse models of LPS-induced peritonitis, dextran sodium sulfate-induced colitis, and high-fat diet-induced non-alcoholic steatohepatitis were used to test whether butein has protective effects on these NLRP3-driven diseases. Results Butein blocks NLRP3 inflammasome activation in mouse macrophages by inhibiting ASC oligomerization, suppressing reactive oxygen species production, and upregulating the expression of the antioxidant pathway nuclear factor erythroid 2-related factor 2 (Nrf2). Importantly, in vivo experiments demonstrated that butein administration has a significant protective effect on the mouse models of LPS-induced peritonitis, dextran sodium sulfate-induced colitis, and high-fat diet-induced non-alcoholic steatohepatitis. Conclusion Our study illustrates the connotation of homotherapy for heteropathy, i.e., the application of butein to broaden therapeutic approaches and treat multiple inflammatory diseases driven by NLRP3.
... ROS play a multifaceted role in regulating NLRP3 inflammasome activation through various pathways. 42 In our prior work, we demonstrated that NADPH oxidase-derived ROS contribute to the positive regulation of the priming signal of the NLRP3 inflammasome. Inhibiting NADPH oxidase or scavenging ROS effectively reduced the LPS-induced expression of NLRP3 and proIL-1β in macrophages. ...
Article
Full-text available
Purpose The NOD-like receptor family pyrin domain-containing 3 (NLRP3) inflammasome, crucial in infectious and inflammatory diseases by regulating IL-1β, presents a target for disease management. Neisseria gonorrhoeae causes gonorrhea in over 87 million people annually, with previous research revealing NLRP3 inflammasome activation in infected macrophages. No natural products have been reported to counteract this activation. Exploring honokiol, a phenolic compound from Chinese herbal medicine, we investigated its impact on NLRP3 inflammasome activation in N. gonorrhoeae-infected macrophages. Methods Honokiol’s impact on the protein expression of pro-inflammatory mediators was analyzed using ELISA and Western blotting. The generation of intracellular H2O2 and mitochondrial reactive oxygen species (ROS) was detected through specific fluorescent probes (CM-H2DCFDA and MitoSOX, respectively) and analyzed by flow cytometry. Mitochondrial membrane integrity was assessed using specific fluorescent probes (MitoTracker and DiOC2(3)) and analyzed by flow cytometry. Additionally, the effect of honokiol on the viability of N. gonorrhoeae was examined through an in vitro colony-forming units assay. Results Honokiol effectively inhibits caspase-1, caspase-11 and GSDMD activation and reduces the extracellular release of IL-1β, NLRP3, and apoptosis-associated speck-like protein containing a caspase recruitment domain (ASC) in N. gonorrhoeae-infected macrophages. Detailed investigations have demonstrated that honokiol lowers the production of H2O2 and the phosphorylation of ERK1/2 in N. gonorrhoeae-infected macrophages. Importantly, the phosphorylation of JNK1/2 and p38 and the activation of NF-κB remain unaffected. Moreover, honokiol reduces the N. gonorrhoeae-mediated generation of reactive oxygen species within the mitochondria, preserving their integrity. Additionally, honokiol suppresses the expression of the pro-inflammatory mediator IL-6 and inducible nitric oxide synthase induced by N. gonorrhoeae independently of NLRP3. Impressively, honokiol exhibits in vitro anti-gonococcal activity against N. gonorrhoeae. Conclusion Honokiol inhibits the NLRP3 inflammasome in N. gonorrhoeae-infected macrophages and holds great promise for further development as an active ingredient in the prevention and treatment of symptoms associated with gonorrhea.
... The production of ROS may damage mitochondria and induce the activation of the NLRP3 inflammasome complex [62][63][64][65]. Specifically, caspase-1, as a downstream protein, is activated by NLRP3 and further induces the gasdermin D-dependent pyroptosis pathway [66][67][68]. ...
Article
Full-text available
The endothelial barrier plays a critical role in immune defense against bacterial infection. Efficient interactions between neutrophils and endothelial cells facilitate the activation of both cell types. However, neutrophil activation can have dual effects, promoting bacterial clearance on one hand while triggering inflammation on the other. In this review, we provide a detailed overview of the cellular defense progression when neutrophils encounter bacteria, focusing specifically on neutrophil–endothelial interactions and endothelial activation or dysfunction. By elucidating the underlying mechanisms of inflammatory pathways, potential therapeutic targets for inflammation caused by endothelial dysfunction may be identified. Overall, our comprehensive understanding of neutrophil–endothelial interactions in modulating innate immunity provides deeper insights into therapeutic strategies for infectious diseases and further promotes the development of antibacterial and anti-inflammatory drugs.
... ROS production is important for pyroptosis, which mediates the activation of NLRP3 inflammasome. 30 Mounting evidence showed that HUVECs generate excessive amounts of ROS in the presence of LPS-ATP that contributed in oxidative stress. Hence, we interrogated the influence of Pue on LPS-ATP-primed ROS generation. ...
Article
Full-text available
The occurrence and development of diabetic vascular diseases are closely linked to inflammation‐induced endothelial dysfunction. Puerarin (Pue), the primary component of Pueraria lobata, possesses potent anti‐inflammatory properties. However, its vasoprotective role remains elusive. Therefore, we investigated whether Pue can effectively protect against vascular damage induced by diabetes. In the study, Pue ameliorated lipopolysaccharide‐adenosine triphosphate (LPS‐ATP) or HG‐primed cytotoxicity and apoptosis, while inhibited reactive oxygen species (ROS)‐mediated NLR family pyrin domain containing 3 (NLRP3) inflammasome in HUVECs, as evidenced by significantly decreased ROS level, NOX4, Caspase‐1 activity and expression of NLRP3, GSDMD, cleaved caspase‐1, IL‐1β and IL‐18. Meanwhile, ROS inducer CoCI2 efficiently weakened the effects of Pue against LPS‐ATP‐primed pyroptosis. In addition, NLRP3 knockdown notably enhanced Pue's ability to suppress pyroptosis in LPS‐ATP‐primed HUVECs, whereas overexpression of NLRP3 reversed the inhibitory effects of Pue. Furthermore, Pue inhibited the expression of ROS and NLRP3 inflammasome‐associated proteins on the aorta in type 2 diabetes mellitus rats. Our findings indicated that Pue might ameliorate LPS‐ATP or HG‐primed damage in HUVECs by inactivating the ROS‐NLRP3 signalling pathway.
Article
Full-text available
Background With the worsening of the greenhouse effect, the correlation between the damp-heat environment (DH) and the incidence of various diseases has gained increasing attention. Previous studies have demonstrated that DH can lead to intestinal disorders, enteritis, and an up-regulation of NOD-like receptor protein 3 (NLRP3). However, the mechanism of NLRP3 in this process remains unclear. Methods We established a DH animal model to observe the impact of a high temperature and humidity environment on the mice. We sequenced the 16S rRNA of mouse feces, and the RNA transcriptome of intestinal tissue, as well as the levels of cytokines including interferon (IFN)-γ and interleukin (IL)-4 in serum. Results Our results indicate that the intestinal macrophage infiltration and the expression of inflammatory genes were increased in mice challenged with DH for 14 days, while the M2 macrophages were decreased in Nlrp3 -/- mice. The alpha diversity of intestinal bacteria in Nlrp3 -/- mice was significantly higher than that in control mice, including an up-regulation of the Firmicutes/Bacteroidetes ratio. Transcriptomic analysis revealed 307 differentially expressed genes were decreased in Nlrp3 -/- mice compared with control mice, which was related to humoral immune response, complement activation, phagocytic recognition, malaria and inflammatory bowel disease. The ratio of IFN-γ/IL-4 was decreased in control mice but increased in Nlrp3 -/- mice. Conclusions Our study found that the inflammation induced by DH promotes Th2-mediated immunity via NLRP3, which is closely related to the disruption of intestinal flora.
Article
Full-text available
Endothelial cells have emerged as key players in SARS-CoV-2 infection and COVID-19 inflammatory pathologies. Dysfunctional endothelial cells can promote chronic inflammation and disease processes like thrombosis, atherosclerosis, and lung injury. In endothelial cells, mitochondria regulate these inflammatory pathways via redox signaling, which is primarily achieved through mitochondrial reactive oxygen species (mtROS). Excess mtROS causes oxidative stress that can initiate and exacerbate senescence, a state that promotes inflammation and chronic endothelial dysfunction. Oxidative stress can also activate feedback loops that perpetuate mitochondrial dysfunction, mtROS overproduction, and inflammation. In this review, we provide an overview of phenotypes mediated by mtROS in endothelial cells – such as mitochondrial dysfunction, inflammation, and senescence – as well as how these chronic states may be initiated by SARS-CoV-2 infection of endothelial cells. We also propose that SARS-CoV-2 activates mtROS-mediated feedback loops that cause long-term changes in host redox status and endothelial function, promoting cardiovascular disease and lung injury after recovery from COVID-19. Finally, we discuss the implications of these proposed pathways on long-term vascular health and potential treatments to address these chronic conditions.
Article
Full-text available
One of the major mechanisms of action of chemo-radiation is to induce cellular senescence, which exerts crucial roles in age-related pathology. The concept of senescence is evolved, and the novel understanding of senescence-associated reprogramming/stemness has emerged. This new concept emphasizes senescence as not only cell cycle arrest but describes that subsets of senescent cells induced by chemotherapy can re-enter cell cycles, proliferate rapidly, and acquire “stemness” status. Cancer therapeutics, including chemo-radiation triggers toxicity effects through damaging mitochondria, primarily through the upregulation of mtROS production leading to subsequent mtDNA and telomeric DNA damage elicitng DNA damage responses (DDR). The ultimate goal of this review is to highlight the new concept of senescence-associated stemness that is induced by cancer treatment and its adverse effects on the vascular system. We will describe how chemo-radiation exerts toxicity effects by simultaneously producing reactive oxygen species in mitochondria and promoting DDR in the nucleus. We discuss the potential of clinical targeting poly (ADP-ribose) polymerase which might prevent downstream mitochondrial dysfunction and confer protection to cancer survivors. Overall we emphasize the importance of recognizing the consequences of cardio-toxic effects of several cancer treatments and therefore developing personalized therapeutic approaches to screen for inflammatory and cardiac testing for better patient survival.
Article
Full-text available
Inflammasomes are a class of cytosolic protein complexes. They act as cytosolic innate immune signal receptors to sense pathogens and initiate inflammatory responses under physiological and pathological conditions. The NLR-family pyrin domain-containing protein 3 (NLRP3) inflammasome is the most characteristic multimeric protein complex. Its activation triggers the cleavage of pro-interleukin (IL)-1β and pro-IL-18, which are mediated by caspase-1, and secretes mature forms of these mediators from cells to promote the further inflammatory process and oxidative stress. Simultaneously, cells undergo pro-inflammatory programmed cell death, termed pyroptosis. The danger signals for activating NLRP3 inflammasome are very extensive, especially reactive oxygen species (ROS), which act as an intermediate trigger to activate NLRP3 inflammasome, exacerbating subsequent inflammatory cascades and cell damage. Vascular endothelium at the site of inflammation is actively involved in the regulation of inflammation progression with important implications for cardiovascular homeostasis as a dynamically adaptable interface. Endothelial dysfunction is a hallmark and predictor for cardiovascular ailments or adverse cardiovascular events, such as coronary artery disease, diabetes mellitus, hypertension, and hypercholesterolemia. The loss of proper endothelial function may lead to tissue swelling, chronic inflammation, and the formation of thrombi. As such, elimination of endothelial cell inflammation or activation is of clinical relevance. In this review, we provided a comprehensive perspective on the pivotal role of NLRP3 inflammasome activation in aggravating oxidative stress and endothelial dysfunction and the possible underlying mechanisms. Furthermore, we highlighted the contribution of noncoding RNAs to NLRP3 inflammasome activation-associated endothelial dysfunction, and outlined potential clinical drugs targeting NLRP3 inflammasome involved in endothelial dysfunction. Collectively, this summary provides recent developments and perspectives on how NLRP3 inflammasome interferes with endothelial dysfunction and the potential research value of NLRP3 inflammasome as a potential mediator of endothelial dysfunction.
Article
Full-text available
Background: In developed countries, the extension of human life is increasingly accompanied by a progressive increase in neurodegenerative diseases, most of which do not yet have effective therapy but only symptomatic treatments. In recent years, plant polyphenols have aroused considerable interest in the scientific community. The mechanisms currently hypothesized for the pathogenesis of Parkinson's disease (PD) are neuroinflammation, oxidative stress and apoptosis. Hydroxytyrosol (HT), the main component of Hidrox® (HD), has been shown to have some of the highest free radical evacuation and anti-inflammatory activities. Here we wanted to study the role of HD on the neurobiological and behavioral alterations induced by rotenone. Methods: A study was conducted in which mice received HD (10 mg/kg, i.p.) concomitantly with rotenone (5 mg/kg, o.s.) for 28 days. Results: Locomotor activity, catalepsy, histological damage and several characteristic markers of the PD, such as the dopamine transporter (DAT) content, tyrosine hydroxylase (TH) and accumulation of α-synuclein, have been evaluated. Moreover, we observed the effects of HD on oxidative stress, neuroinflammation, apoptosis and inflammasomes. Taken together, the results obtained highlight HD's ability to reduce the loss of dopaminergic neurons and the damage associated with it by counteracting the three main mechanisms of PD pathogenesis. Conclusion: HD is subject to fewer regulations than traditional drugs to improve patients' brain health and could represent a promising nutraceutical choice to prevent PD.
Article
Full-text available
Accumulation of senescent cells has a causative role in the pathology of age-related disorders including atherosclerosis (AS) and cardiovascular diseases (CVDs). However, the concept of senescence is now drastically changing, and the new concept of senescence-associated reprogramming/stemness has emerged, suggesting that senescence is not merely related to “cell cycle arrest” or halting various cellular functions. It is well known that disturbed flow (D-flow) accelerates pre-mature aging and may play a role in the development of AS. We will discuss in this review that pre-mature aging induced by d-flow is not comparable to time-dependent aging, particularly with a focus on the possible involvement of senescence-associated secretory phenotype (SASP) in senescence-associated reprogramming/stemness, or increasing cell numbers. We will also present our outlook of nicotinamide adenine dinucleotides (NAD)⁺ deficiency-induced mitochondrial reactive oxygen species (mtROS) in evoking SASP by activating DNA damage response (DDR). MtROS plays a key role in developing cross-talk between nuclear-mitochondria, SASP, and ultimately atherosclerosis formation. Although senescence induced by time and various stresses including D-flow is a classic concept, we wish the readers will see the undergoing Copernican-like change in this concept, as well as the difference between pre-mature aging induced by D-flow and time-dependent aging.
Article
Full-text available
Inflammation is a protective reaction activated in response to detrimental stimuli, such as dead cells, irritants or pathogens, by the evolutionarily conserved immune system and is regulated by the host. The inflammasomes are recognized as innate immune system sensors and receptors that manage the activation of caspase-1 and stimulate inflammation response. They have been associated with several inflammatory disorders. The NLRP3 inflammasome is the most well characterized. It is so called because NLRP3 belongs to the family of nucleotide-binding and oligomerization domain-like receptors (NLRs). Recent evidence has greatly improved our understanding of the mechanisms by which the NLRP3 inflammasome is activated. Additionally, increasing data in animal models, supported by human studies, strongly implicate the involvement of the inflammasome in the initiation or progression of disorders with a high impact on public health, such as metabolic pathologies (obesity, type 2 diabetes, atherosclerosis), cardiovascular diseases (ischemic and non-ischemic heart disease), inflammatory issues (liver diseases, inflammatory bowel diseases, gut microbiome, rheumatoid arthritis) and neurologic disorders (Parkinson’s disease, Alzheimer’s disease, multiple sclerosis, amyotrophic lateral sclerosis and other neurological disorders), compared to other molecular platforms. This review will provide a focus on the available knowledge about the NLRP3 inflammasome role in these pathologies and describe the balance between the activation of the harmful and beneficial inflammasome so that new therapies can be created for patients with these diseases.
Article
Full-text available
Inflammasomes are cytoplasmic multiprotein complexes comprising a sensor protein, inflammatory caspases, and in some but not all cases an adapter protein connecting the two. They can be activated by a repertoire of endogenous and exogenous stimuli, leading to enzymatic activation of canonical caspase-1, noncanonical caspase-11 (or the equivalent caspase-4 and caspase-5 in humans) or caspase-8, resulting in secretion of IL-1β and IL-18, as well as apoptotic and pyroptotic cell death. Appropriate inflammasome activation is vital for the host to cope with foreign pathogens or tissue damage, while aberrant inflammasome activation can cause uncontrolled tissue responses that may contribute to various diseases, including autoinflammatory disorders, cardiometabolic diseases, cancer and neurodegenerative diseases. Therefore, it is imperative to maintain a fine balance between inflammasome activation and inhibition, which requires a fine-tuned regulation of inflammasome assembly and effector function. Recently, a growing body of studies have been focusing on delineating the structural and molecular mechanisms underlying the regulation of inflammasome signaling. In the present review, we summarize the most recent advances and remaining challenges in understanding the ordered inflammasome assembly and activation upon sensing of diverse stimuli, as well as the tight regulations of these processes. Furthermore, we review recent progress and challenges in translating inflammasome research into therapeutic tools, aimed at modifying inflammasome-regulated human diseases.
Article
Inhibiting the NLRP3 inflammasome mediates inflammation in an extensive number of preclinical models. As excitement in this field has grown, several companies have recently initiated testing of direct NLRP3 inhibitors in the clinic. At the same time, the NLRP3 inflammasome is part of a larger pro-inflammatory pathway, whose modulation is also being explored. Multiple targets in this pathway are already impinged upon by molecules that have been through clinical trials. These data, informed by the growing mechanistic understanding of the NLRP3 inflammasome in the preclinical space, provide a rich backdrop to assess the current state of the field. Here we explore attempts to inhibit the NLRP3 inflammasome in light of clinical and preclinical data around efficacy and safety.
Article
The NLRP3 (NOD‐, LRR‐, and pyrin domain‐containing protein 3) inflammasome is an immunological sensor that detects a wide range of microbial‐ and host‐derived signals. Inflammasome activation results in the release of the potent pro‐inflammatory cytokines IL‐1β and IL‐18 and triggers a form of inflammatory cell death known as pyroptosis. Excessive NLRP3 activity is associated with the pathogenesis of a wide range of inflammatory diseases, thus NLRP3 activation mechanisms are an area of intensive research. NLRP3 inflammasome activation is a tightly regulated process that requires both priming and activation signals. In particular, recent research has highlighted the highly complex nature of the priming step, which involves transcriptional and posttranslational mechanisms, and numerous protein binding partners. This review will describe the current understanding of NLRP3 priming and will discuss the potential opportunities for targeting this process therapeutically to treat NLRP3‐associated diseases. Describes the complex NLRP3 priming process and how it might be targeted to treat NLRP3‐associated diseases.
Article
Forty years after its naming, interleukin-1 (IL-1) is experiencing a renaissance brought on by the growing understanding of its context-dependent roles and advances in the clinic. Recent studies have identified important roles for members of the IL-1 family-IL-18, IL-33, IL-36, IL-37, and IL-38-in inflammation and immunity. Here, we review the complex functions of IL-1 family members in the orchestration of innate and adaptive immune responses and their diversity and plasticity. We discuss the varied roles of IL-1 family members in immune homeostasis and their contribution to pathologies, including autoimmunity and auto-inflammation, dysmetabolism, cardiovascular disorders, and cancer. The trans-disease therapeutic activity of anti-IL-1 strategies argues for immunity and inflammation as a metanarrative of modern medicine.