You are on page 1of 4

616

FORUM

Eur. J. Immunol. 2010. 40: 595653

Activation mechanisms
Signaling by ROS drives inflammasome activation
Fabio Martinon
Department of Immunology and Infectious Diseases, Harvard School of Public Health,
Boston, MA, USA

Inflammasomes are innate immune


signaling pathways that sense pathogens and injury to direct the proteolytic maturation of inflammatory
cytokines such as IL-1b and IL-18.
Among inflammasomes, the NLRP3/
NALP3 inflammasome is the most
studied. However, little is known on
the molecular mechanisms that
mediate its assembly and activation.
Recent findings suggest that ROS are
produced by NLRP3/NALP3 activators
and are essential secondary messengers signaling NLRP3/NALP3 inflammasome activation.

Redox signaling and oxidative


stress
ROS are free radicals that contain the
oxygen atom and include hydrogen
peroxide (H2O2), superoxide anion
(O2!" ) and hydroxyl radical (OH!).
These molecules are highly reactive
(oxidizing/electron-capturing) due to
the presence of unpaired valence shell
electron. ROS mainly originate as a
byproduct of oxygen metabolism in the
electron transport chain within the
mitochondria (Fig. 1). ROS are also
generated by the activity of cellular
enzymes such as NADPH oxidases,
xanthine oxidoreductases, lipoxygenases
and cyclooxygenases [1]. Cellular
production of ROS regulates several
important physiological responses, such
as oxygen sensing, angiogenesis, control
of vascular tone, and regulation of cell
growth, differentiation and migration.
While ROS is also important for cell
signaling (a phenomenon known as
redox signaling), sustained ROS production can cause cellular damage. To cope
with this stress, several enzymes display-

DOI 10.1002/eji.200940168

ing anti-oxidant activities, including


thioredoxin (TRX), superoxide dismutases, glutathione peroxidases and catalase, are involved in neutralizing ROS.
The imbalance between the formation of
ROS and the ability to detoxify these
oxidizing radicals can produce a cellular
state known as oxidative stress [2]. ROSmediated oxidative stress plays an
important role in pathological processes
such as aging, hypertension, atherosclerosis, cancer, ischemia, neurodegenerative diseases and diabetes [1].
Production of reactive ROS is crucial
to the regulation of innate immune
responses. In plants for example,
pathogen recognition generates ROS in
an NADPH oxidase-dependent manner
to cause the oxidative burst leading to
the hypersensitive response [3]. Similarly, during inflammation and immune
responses in vertebrates, activated
phagocytic cells such as neutrophils
generate a ROS-dependent respiratory
burst that directs toxicity towards
invading microbes [4]. ROS is also
involved in signaling injury to the
immune system. Beyond its antiseptic
function, release of ROS (H2O2) by
damaged tissues can form a decreasing
concentration gradient that directs
leukocytes recruitment at the site of
tissue injury, demonstrating that ROS
can orchestrate inflammatory responses
in tissues [5]. Redox signaling is also
important in the signaling pathways
engaged by various inflammatory
conditions. ROS production by the PRR,
TLR, regulates activation of redoxregulated transcription factors (NF-kB
and AP-1) and cytokines production
[6, 7]. Recently, ROS has been
proposed to play an important role in
the activation of the NLRP3/NALP3
inflammasome [8, 9].

& 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ROS is required for NLRP3/NALP3


Inflammasome activation
The inflammasome is a cytosolic molecular complex that once activated has
an enzymatic activity mediated by the
recruitment and activation of caspase-1.
The inflammasome senses pathogens
and stress or danger signals to promote
the maturation of cytokines such as
IL-1b. The release of active IL-1b
engages IL-1 receptor-harboring cells
and promotes inflammatory responses
[10]. Although the activation of
caspase-1 and the maturation of IL-1b
is common to virtually all kinds of
inflammasomes, the scaffolding unit
involved in sensing pathogens or
danger signals may vary [9]. The NLR
protein NLRP3/NALP3 forms the most
studied inflammasome. Upon activation, NLRP3/NALP3 recruits the adaptor ASC and the enzyme caspase-1 to
form the NLRP3/NALP3 inflammasome.
No NLRP3/NALP3 activators have been
shown to directly interact and activate
NLRP3/NALP3, suggesting that NLRP3/
NALP3 may sense these signals indirectly. Interestingly, most identified
NLRP3/NALP3 activators also trigger
ROS production. Moreover, the use of
antioxidants has been shown to inhibit
NLRP3/NALP3 inflammasome activation, suggesting that redox signaling or
oxidative stress is involved in NLRP3/
NALP3 activation.
Extracellular ATP is an inflammatory
signal that has been implicated in
innate immunity in both plants and
animals [11]. In mammals, extracellular ATP binds to P2X7 receptors
and activates the NLRP3/NALP3
inflammasome [12]. Treatment of
macrophages with ATP results in the
rapid production of ROS and the use of

www.eji-journal.eu

FORUM

Eur. J. Immunol. 2010. 40: 595653

vators, such as the toxin nigericin, UV


light and skin sensitizers (e.g. dinitrochlorobenzene) activate a cellular redox
imbalance required for inflammasome
formation [14, 22, 23]. ROS has also
been implicated in NLRP3/NALP3 activation by the malaria pathogenic crystal, hemozoin, the influenza virus and
the yeast Candida albicans [2426].

How do NLRP3/NALP3 activators


promote ROS generation?

Figure 1. Examples of ROS generating pathways. (A) During respiration 12% of the
oxygen is partially reduced to O2 " , which
can be converted H2O2 and OH . The major
sites of for production in the mitochondrial
respiratory chain are at complexes I and III.
Complex I accepts electrons from NADH;
these electrons move down an electrochemical gradient through ubiquinone (Q cycle) to
complex III, from complex III to cytochrome c
(C) and from C to complex IV that use the
electrons to reduce molecular oxygen to
water. The mechanisms involved in
generation of O2 " by complex I are poorly
understood. Complex III generate O2 " by
auto-oxidation of ubisemiquinone generated
during the Q cycle, (IM, inner mitochondrial
membrane). (B) NADPH oxidases such as the
NOX2 complex transport electrons across
biological membranes to reduce oxygen to
superoxide. The activation of NOX2 occurs
through a complex series of protein/protein
interactions. Phosphorylation of p47phox leads
to a conformational change allowing its
interaction with p22phox. The localization of
p47phox to the membrane brings p67phox into
contact with NOX2 and also brings the small
subunit p40phox to the complex. Finally, the
GTPase Rac interacts with NOX2. Once
assembled, the complex is active and generates superoxide by transferring an electron
from NADPH in the cytosol to oxygen on the
luminal or extracellular space.

the broad spectrum NADPH oxidase


inhibitor, diphenyleneiodonium (DPI)
inhibits ATP-mediated caspase-1 activation [13, 14]. The NLRP3/NALP3
activating particulate elements uric acid
crystals, alum and particulate metals
have been shown to induce ROS
production [1518]. Similarly, ROS is
detectable quickly upon exposure of
macrophages to silica or asbestos
[16, 1921] Other NLRP3/NALP3 acti-

Various pathways have been proposed


to mediate ROS production by NLRP3/
NALP3 activators; however, the general
picture of how NLRP3/NALP3 activators trigger ROS is still unclear. Potassium efflux and decrease in cytosolic
potassium concentration are the most
striking features associated with
NLRP3/NALP3 activators [27]. Interestingly, potassium efflux has been linked
to ROS production at the membrane in
plants [28]. Moreover, potassium efflux
has been shown to trigger ROS production in human granulocytes [29]. It is
therefore tempting to speculate that
potassium efflux by NLRP3/NALP3 activators could be involved in ROS generation.
Some NLRP3/NALP3 activators such
as uric acid crystals, alum, asbestos and
silica are large particulate elements that
can induce the so-called frustrated
phagocytosis at the cell surface.
Frustrated phagocytosis has been associated with ROS production; however,
the mechanism by which this occurs is
unclear [30, 31]. Frustrated phagocytosis may not be the only mechanism
used by macrophages to sense pathogenic particles. Evidence has demonstrated that uric acid crystals can be
phagocytosed. Ultrastructure studies of
uric acid crystal-containing phagolysosomes show a disrupted membrane and
possibly release of part of their content
in the cytoplasm [32]. In line with these
early observations, silica crystals and
alum trigger damage and rupture of the
lysosome [33], as described by the
companion Viewpoint article [34].
Importantly, the release of cathepsin B
by damaged lysosomes has been
proposed to mediate inflammasome

& 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

activation [33]. It is unclear whether


this mechanism works in parallel to the
ionic imbalance and oxidative stress
pathway. It is also possible that lysosomal damage and cathepsin B release act
upstream of ROS production. In line
with this model, cathepsin B has been
shown to promote ROS production in
hepatocytes and neurones [35, 36].
Multiple lines of evidence suggest
that ROS poduction by NLRP3/NALP3
activators involves NADPH-oxidases
(NOX). NOX are a family of transmembrane enzymes that generate ROS by
carrying electrons across biological
membranes from a cytosolic electron
donor (such as NADPH) to an electron
acceptor (oxygen) in the extracellular or
luminal space [37]. The observation
that NOX inhibitors such as DPI or
(2R,4R)-4-aminopyrrolidine-2,4-dicarboxylate inhibits inflammasome activation by virtually all NLRP3/NALP3
activators identified so far suggests that
NOX are involved in ROS production.
This has also been demonstrated in vivo.
Indeed, DPI inhibits caspase-1 mediated
IL-18 activation in mice undergoing
physical stress [38].
NOX inhibitors may have additional
targets. DPI can exert inhibitory effect
on mitochondrial ROS production in
addition to NOX [39]. However, in line
with the possibility that NOX are
involved in inflammasome activation,
extracellular ATP has been shown to
trigger translocation of cytosolic NOX
components
(p47phox,
p40phox,
p67phox and p21rac) onto membranebound NOX2, forming an active
macromolecular complex [14]. Indeed,
NOX2 deficiency impairs ATP-mediated
ROS production by macrophages,
suggesting that NOX2 may be involved
in ATP-mediated NLRP3/NALP3 activation [40]. On the contrary, NOX2-deficient macrophages have no defect in
inflammasome activation upon stimulation with other NLRP3/NALP3
agonists including uric acid crystals and
silica [33] while knockdown of
p22phox, in the monocytic cell line
THP1 impairs inflammasome activation
by hemozoin, silica uric acid crystal and
asbestos [16, 41]. Because p22phox
deficiency affects several NOX including
NOX1, NOX2, NOX3 and NOX4 [37], it

www.eji-journal.eu

617

618

FORUM

is possible that multiple NOX mediate


ROS production to trigger NLRP3/
NALP3 inflammasome assembly.
Overall, many studies using antioxidants support a model in which ROS
production by NLRP3/NALP3 agonists
drive inflammasome assembly. However,
the mechanisms of production and the
nature of ROS involved in inflammasome
activation are unknown. Future work
should focus on characterizing how
frustrated phagocytosis, cathepsin B,
potassium efflux and NOX may synergize
and contribute to ROS production and
inflammasome activation.

How does ROS trigger NLRP3/


NALP3 activation?
ROS production by H2O2 activates the
inflammasome [16, 42], furthermore,
knockdown of TRX, a cellular antioxidant protein, enhances IL-1b activation by silica, uric acid crystal and
asbestos [41]. These findings suggest
that oxidative stress could be sufficient
to trigger NLRP3/NALP3 activation and
lead to interrogate how NLRP3/NALP3
senses ROS. ROS may either directly
trigger inflammasome assembly or be
indirectly sensed through cytoplasmic
proteins that modulate inflammasome
activity. ATP-mediated ROS production
has been shown to stimulate the PI3K
pathway, and pharmacological inhibition of PI3K inhibits ATP-mediated
caspase-1 activation suggesting that
PI3K may be involved in inflammasome
activation downstream of ROS [13].
Recently Ju
rg Tschopps laboratory
identified TXNIP/VDUP1 as an essential
protein that may directly activate
NLRP3/NALP3 upon oxidative stress
[42]. The authors of this study suggest
that, in resting cells, TXNIP/VDUP1
interacts with TRX and is therefore
unable to activate NLRP3/NALP3. Upon
oxidative stress TXNIP is released from
oxidized TRX and in turn directly binds
the leucine-rich region of NLRP3/
NALP3 leading to inflammasome
assembly [42]. Consistent with this
finding, TXNIP/VDUP1-deficient macrophages treated with extracellular ATP
or uric acid crystals have decreased
caspase-1 and IL-1b processing [42].

Eur. J. Immunol. 2010. 40: 595653

This finding provides support for a


model in which TXNIP/VDUP1 and
NLRP3/NALP3 set up a surveillance of
cellular stress, preparing to drive
inflammation in case of excessive stress
or danger signals.

Concluding remarks
In contradiction with classical PRR,
rather than directly recognizing pathogenic elements, NLRP3/NALP3 seems to
detect oxidative stress produced by
pathogenic insults [43] (Fig. 2). This
model shares similarities with the guard
model in plants. Pathogen-mediated
changes in plant cellular physiology
trigger activation of the R genes, a
family of innate immune sensors that
cope with infections and share structural
similarities with NLRP3/NALP3 [44].
Although the mechanisms involved in
inflammasome activation by oxidative
stress are still unclear, ROS is emerging

as the central and common element


regulating NLRP3/NALP3 activation. A
fine-tuning of the events underlining
inflammasome activation and inflammation responses by oxidative stress is
likely key to proper immunity and tissue
repair. In line with the role of ROS in
activating NLRP3/NALP3, inhibition of
ROS production in M2 polarized macrophages dampens inflammasome activation [45]. On the other hand, prolonged
oxidative stress can dampen inflammatory mediators [12] including inhibition
of caspase-1 by reversible oxidation and
glutathionylation of redox-sensitive
cysteine residues [46], suggesting that
beyond its role in activating NLRP3/
NALP3, oxidative stress may be part of a
regulatory loop negatively regulating
IL-1b activation.
Oxidative stress and tissue injury are
major hallmarks of numerous pathologies
ranging from diabetes to neurodegenerative disorders. Most of theses pathologies have an inflammatory component

Figure 2. Model of NLRP3/NALP3 inflammasome assembly and activation. Multiple NLRP3/


NALP3 inflammasome activators such as extracellular ATP and particulate elements trigger ROS
production. Possible pathways involved in ROS production include potassium efflux, frustrated
phagocytosis, phagolysosomes disruption, Cathepsin B release and NOX activation. Oxidative
stress triggers inflammasome-activating signals such as PI3K and TXNIP release from oxidized
TRX. Binding of TXNIP to NLRP3/NALP3 promotes assembly and oligomerization of the
inflammasome. The recruitment of the adaptor ASC and the enzyme caspase-1 to the
inflammasome are crucial for its proIL-1b cleaving activity.

& 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eji-journal.eu

FORUM

Eur. J. Immunol. 2010. 40: 595653

[47]. Dissecting the possible involvement


of the inflammasome in such pathologies
and identifying how oxidative stress
regulates NLRP3/NALP3 activation and
IL-1b activity will likely shed some new
light on these pathologies.

Acknowledgements:
The
author
is
supported by a Human Frontier Science
Program long-term fellowship.
Conflict of interest: The author declares no
financial or commercial conflict of interest.
ge, W., Physiol. Rev. 2002. 82: 4795.
1 Dro
2 Valko, M. et al., Int. J. Biochem. Cell Biol.
2007. 39: 4484.
3 Ma, W. and Berkowitz, G. A., Cell
Microbiol. 2007. 9: 25712585.
4 Bogdan, C. et al., Curr. Opin. Immunol.
2000. 12: 6476.
5 Niethammer, P. et al., Nature 2009.
459: 996999.
6 Ogier-Denis, E. et al., Semin. Immunopathol. 2008. 30: 291300.
7 Iriti, M. and Faoro, F., Mycopathologia
2007. 164: 5764.
8 Martinon, F. et al., Annu. Rev. Immunol.
2009. 27: 229265.
9 Bryant, C. and Fitzgerald, K. A., Trends
Cell Biol. 2009. 19: 455464.
10 Martinon, F. et al., Mol. Cell 2002. 10:
417426.
11 Chivasa, S. et al., Proteomics 2010. 10:
235244.
12 Carta, S. et al., J. Leukoc. Biol. 2009:
17.
13 Cruz, C. M. et al., J. Biol. Chem. 2007.
282: 28712879.
14 Hewinson, J. et al., J. Immunol. 2008.
180: 84108420.
15 Eisenbarth, S. C. et al., Nature 2008.
453: 11221126.

16 Dostert, C. et al., Science 2008. 320:


674677.
17 Sautin, Y. Y. et al., Am. J. Physiol. Cell
Physiol. 2007. 293: C584C596.
18 Caicedo, M. et al., J. Orthop. Res. 2009
27: 847854.
19 Cassel, S. L. et al., Proc. Natl. Acad. Sci.
USA 2008. 105: 90359040.
20 Fubini, B. and Hubbard, A., Free Radic.
Biol. Med. 2003. 34: 15071516.
21 Simeonova, P. P. and Luster, M. I., Am.
J. Respir. Cell Mol. Biol. 1995. 12:
676683.
22 Feldmeyer, L. et al., Curr. Biol. 2007.
17: 11401145.
23 Jin, G.-H. et al., Radiat. Environ. Biophys.
2007. 46: 6168.
24 Jaramillo, M. et al., J. Immunol. 2005.
174: 475484.
25 Gross, O. et al., Nature 2009 459:
433436.
26 Allen, I. C. et al., Immunity 2009 30:
556565.
trilli, V. et al., Cell Death Differ. 2007.
27 Pe
14: 15831589.
28 Bolwell, G. P., Curr. Opin. Plant Biol.
1999. 2: 287294.
29 Fay, A. J. et al., Proc. Natl. Acad. Sci. USA
2006. 103: 1754817553.
30 ONeill, L. A. J., Science 2008. 320:
619620.
31 Bergstrand, H., Agents Actions Suppl.
1990. 30: 199211.
32 Hoffstein, S. and Weissmann, G.,
Arthritis Rheum. 1975. 18: 153165.
33 Hornung, V. et al., Nat. Immunol. 2008.
9: 847856.
34 Hornung, V. and Latz, E., Eur. J.
Immunol 2010. 40: 620623.
35 Windelborn, J. A. and Lipton, P., J.
Neurochem. 2008. 106: 5669.
36 Li, Z. et al., Hepatology 2008. 47:
14951503.
37 Bedard, K. and Krause, K.-H., Physiol.
Rev. 2007. 87: 245313.

& 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

38 Sekiyama, A. et al., Immunity 2005. 22:


669677.
39 Aldieri, E. et al., Curr. Drug Metab. 2008.
9: 686696.
40 Moore, S. F. and Mackenzie, A. B., J.
Immunol. 2009. 183: 33023308.
41 Dostert, C. et al., PLoS ONE 2009. 4:
e6510.
42 Zhou, R. et al., Nat. Immunol. 2010.
Published online: 20 December 2009
doi:10.1038/ni.1831.
43 Martinon, F., J. Leukoc. Biol. 2008. 83:
507511.
44 Schneider, D. S., Cell 2002. 109:
537540.
45 Pelegrin, P. and Surprenant, A., EMBO
J. 2009. 28: 21142127.
46 Meissner, F. et al., Nat. Immunol. 2008.
9: 866872.
47 Medzhitov, R., Nature 2008. 454:
428435.

Correspondence: Dr. Fabio Martinon,


Department of Immunology and Infectious
Diseases, Harvard School of Public Health,
651 Huntington Ave, Boston, MA 02115,
USA
Fax: 11-617-432-0084
e-mail: fmartino@hsph.harvard.edu
Received: 18/11/2009
Revised: 21/12/2009
Accepted: 8/1/2010
Key words: Danger signals ! Inflammasome
! Inflammation ! Oxidative stress
Abbreviations: DPI: diphenyleneiodonium !
NOX: NADPH-oxidase ! TRX: thioredoxin
See accompanying Viewpoint:
http://dx.doi.org/10.1002/eji.200940185

www.eji-journal.eu

619

You might also like