ArticlePDF AvailableLiterature Review

AIM2 inflammasome in infection, cancer, and autoimmunity: Role in DNA sensing, inflammation, and innate immunity

Authors:

Abstract and Figures

Recognition of DNA by the cell is an important immunological signature that marks the initiation of an innate immune response. AIM2 is a cytoplasmic sensor that recognizes dsDNA of microbial or host origin. Upon binding to DNA, AIM2 assembles a multi-protein complex called the inflammasome, which drives pyroptosis and proteolytic cleavage of the pro-inflammatory cytokines pro-IL-1β and pro-IL-18. Release of microbial DNA into the cytoplasm during infection by Francisella, Listeria, Mycobacterium, mouse cytomegalovirus, vaccinia virus, Aspergillus and Plasmodium species leads to activation of the AIM2 inflammasome. In contrast, inappropriate recognition of cytoplasmic self-DNA by AIM2 contributes to the development of psoriasis, dermatitis, arthritis and other autoimmune and inflammatory diseases. Inflammasome-independent functions of AIM2 have also been described, including the regulation of the intestinal stem cell proliferation and the gut microbiota ecology in the control of colorectal cancer. In this review we provide an overview of the latest research on AIM2 inflammasome and its role in infection, cancer and autoimmunity. This article is protected by copyright. All rights reserved.
Regulation of the activation of the AIM2 inflammasome. The AIM2 inflammasome is activated by a number of microbial pathogens and dsDNA ligands, including the DNA virus MCMV, the cytosolic bacterium F. novicida, and the dsDNA ligand poly(dA:dT). MCMV infection or transfection of poly(dA:dT) leads to "canonical" activation of the AIM2 inflammasome, which does not require the type I IFN pathway. F. novicida infection activates the AIM2 inflammasome via a "noncanonical" pathway owing to its requirement for type I IFN, analogous to the non-canonical NLRP3 inflammasome pathway. Intracellular F. novicida releases DNA into the cytoplasm to activate the DNA sensors cGAS, STING, and IFI204, which drive transcription of genes encoding type I IFN molecules. It remains unclear why the released DNA is unable to activate AIM2 at this stage, since AIM2 is constitutively expressed in the cell. Type I IFN provides a feedback loop to induce expression of the transcription factor IRF1, which upregulates expression of the IFN-inducible GTPases, including GBP2 and GBP5. GBP2 and GBP5 are recruited to bacterial structures, however, whether they directly target the bacterial membrane or the membrane of intact Francisella-containing vacuole is unclear. Nevertheless, GBPs mediate bacterial killing, resulting in abundant release of bacterial DNA for recognition by AIM2. Assembly of the AIM2 inflammasome induces caspase-1-dependent cleavage of pro-IL-1β and pro-IL-18. Caspase-1 also drives cleavage of the substrate gasdermin D to induce pyroptosis.
… 
Content may be subject to copyright.
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS
DOI: 10.1002/eji.201545839 269
Review
AIM2 inflammasome in infection, cancer,
and autoimmunity: Role in DNA sensing, inflammation,
and innate immunity
Si Ming Man, Rajendra Karki and Thirumala-Devi Kanneganti
Department of Immunology, St. Jude Children’s Research Hospital, Memphis, TN, USA
Recognition of DNA by the cell is an important immunological signature that marks
the initiation of an innate immune response. AIM2 is a cytoplasmic sensor that rec-
ognizes dsDNA of microbial or host origin. Upon binding to DNA, AIM2 assembles a
multiprotein complex called the inflammasome, which drives pyroptosis and proteolytic
cleavage of the proinflammatory cytokines pro-IL-1β and pro-IL-18. Release of microbial
DNA into the cytoplasm during infection by Francisella, Listeria, Mycobacterium, mouse
cytomegalovirus, vaccinia virus, Aspergillus, and Plasmodium species leads to activation
of the AIM2 inflammasome. In contrast, inappropriate recognition of cytoplasmic self-
DNA by AIM2 contributes to the development of psoriasis, dermatitis, arthritis, and other
autoimmune and inflammatory diseases. Inflammasome-independent functions of AIM2
have also been described, including the regulation of the intestinal stem cell proliferation
and the gut microbiota ecology in the control of colorectal cancer. In this review we pro-
vide an overview of the latest research on AIM2 inflammasome and its role in infection,
cancer, and autoimmunity.
Keywords: AIM2 inflammasome
r
Autoimmunity
r
Bacterial/viral infection
r
Cancer
r
DNA sensing
r
Gut microbiota
Introduction
DNA recognition by innate immune receptors triggers a myriad of
immunological responses that are both beneficial and detrimen-
tal to the host. The discovery of Toll-like receptor 9 (TLR9) as a
membrane-associated sensor of bacterial CpG DNA provides evi-
dence for the existence of host receptors that specifically mediate
immune responses to DNA [1]. Translocation of microbial or mam-
malian DNA into the cytoplasm of host cells further induces tran-
scription of genes encoding type I interferon (IFN) molecules and
inflammation independently of TLR9 [2, 3]. Recent advances in
the field h ave identified multiple cytoplasmic DNA sensors that are
responsible for transcriptional activity, including cyclic-GMP-AMP
synthase (cGAS), STING, DDX41, Ku70, LRRFIP1, DNA-dependent
activator of IFN-regulatory factors (IRFs; DAI, also known as
Correspondence: Dr. Thirumala-Devi Kanneganti
e-mail: Thirumala-Devi.Kanneganti@STJUDE.ORG
ZBP1), and IFI16 (reviewed elsewhere [4–6]). Of these DNA sen-
sors, cGAS binds to double-stranded DNA (dsDNA), resulting in a
conformational change in cGAS that allows it to convert ATP and
GTP to a cyclic dinucleotide cyclic-GMP-AMP [7]. Cyclic-GMP-
AMP then binds and activates STING to induce transcription of
genes encoding type I IFN and proinflammatory cytokines via the
transcription factors IRF3 and NF-κB, respectively, (reviewed in
[7]). The molecular basis underlying recognition of DNA by the
other aforementioned cytoplasmic DNA sensors is less understood.
DNA introduced into the cytoplasm also induces IL-1β secre-
tion and pyroptosis [8, 9], and these responses are dependent on
the activity of a cytoplasmic caspase-1-containing complex known
as the inflammasome [10]. In 2009, four groups independently
identified AIM2 as the sensor that triggers inflammasome acti-
vation, pyroptosis, and release of IL-1β and IL-18 in response to
intracellularly delivered dsDNA [11–14].
AIM2 consists of a C-terminal HIN-200 domain, which binds
directly to dsDNA, and an N-terminal pyrin domain (PYD), which
interacts with the PYD of the bipartite PYD-CARD-containing
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
270 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
Figure 1. The molecular basis for the activation of the AIM2 inflam-
masome. The DNA sensor AIM2 is composed of an N-terminal pyrin
domain and a C-terminal HIN-200 domain. The pyrin and HIN-200
domain of AIM2 form an intramolecular complex and are maintained in
an autoinhibitory state. Cytoplasmic dsDNA induces activation of AIM2.
The HIN-200 domain interacts with dsDNA in a sequence-independent
manner, by binding to the sugar-phosphate backbone of dsDNA. The
pyrin domain of AIM2 binds to the pyrin domain of ASC. CARD of ASC
binds the CARD of procaspase-1, forming a macromolecular complex
known as the AIM2 inflammasome. Activated caspase-1 drives cleavage
of pro-IL-1β and pro-IL-18. Caspase-1 also cleaves the substrate gasder-
min D. The N-terminal fragment of gasdermin D induces pyroptosis,
allowing mature IL-1β and IL-18 to be released from the cell.
inflammasome adaptor protein ASC (apoptosis-associated speck-
like protein containing a carboxy-terminal CARD) [10]. The CARD
of ASC binds the CARD of procaspase-1, forming a macromolec-
ular complex fulfilling the basic structural elements of an inflam-
masome (Fig. 1) [10]. AIM2, IFI16, and other pyrin and HIN
domain-containing proteins form the AIM2-like receptor family
[15–17].
AIM2 recognizes dsDNA in a sequence-independent manner;
however, the DNA sequence must be at least 80 base pairs in length
[18]. Elucidation of the crystal structure of AIM2 provided insights
into the activation mechanism of this DNA-sensing inflammasome.
The PYD and HIN-200 domain of AIM2 form an intramolecular
complex and are maintained in an autoinhibitory state during
homeostasis (Fig. 1) [18, 19]. Binding of dsDNA to the HIN-200
domain displaces PYD from the intramolecular complex, liberating
PYD for interaction with ASC [18]. The sugar-phosphate backbone
of dsDNA interacts with the positively charged HIN-200 domain
via electrostatic attraction, allowing sequence-independent recog-
nition of DNA by AIM2. The PYD of AIM2 can also self-oligomerize
to induce the activation of the AIM2 inflammasome [20, 21]. Acti-
vation of AIM2 or a related inflammasome sensor NLRP3 initiates
polymerization of the PYD of ASC [22, 23], ultimately forming
a single and visually distinct inflammasome speck that is readily
observed in primary macrophages and DCs [24–29]. Recent work
has even suggested that the role of PYD of AIM2 is not to maintain
autoinhibition, but to oligomerize and drive filament formation
[30].
Activation of the AIM2 inflammasome and other canonical
inflammasomes results in a type of inflammatory cell death called
pyroptosis [10], which is mediated, in part, by the inflamma-
tory caspase substrate gasdermin D [31, 32] (Fig. 1). Activa-
tion of the AIM2 inflammasome is tightly regulated by the cell
and requires phosphorylation and linear ubiquitination of ASC
[33, 34]. Autophagy can mediate degradation of the AIM2 inflam-
masome to terminate inflammatory responses [35]. Furthermore,
several proteins produced in the cell have the ability to inhibit acti-
vation of the AIM2 inflammasome [14, 36–43]. These include the
PYD-containing proteins POP1 [39] and POP3 [36] in human cells
and the bipartite protein containing two HIN domains, p202, in
mouse cells [14, 40–43]. Here, we summarize the latest research
on the regulation of AIM2 inflammasome, and its role in pathogen
recognition after infection, cancer, and autoimmunity.
AIM2 in the response to infection
Bacterial infection
During infection of a host cell, microbial DNA and other microbe-
associated molecular patterns are released into the cytoplasm,
where they are recognized by cytoplasmic DNA sensors, i.e. cGAS,
STING, or AIM2. AIM2 resides in the cytoplasm and has been
shown to provide immunosurveillance to the pathogenic bacte-
ria Francisella tularensis Live Vaccine Strain (LVS), F. tularensis
subspecies novicida (F. novicida), Listeria monocytogenes, Strep-
tococcus pneumonia, Mycobacterium species, Porphyromonas gingi-
valis, Staphylococcus aureus, Brucella abortus, and Chlamydia muri-
darum (Table 1) [26, 44–61]. F. novicida and F. tularensis LVS are
the only bacterial pathogens known to exclusively activate the
AIM2 inflammasome in mouse macrophages and DCs, whereas
other bacteria have been shown to activate more than one inflam-
masome sensor in mouse and human cells [44–46]. In the human
THP-1 macrophage-like cell line, both AIM2 and NLRP3 contribute
to the activation of the inflammasome in response to F. novicida
and F. tularensis LVS [48]. Interestingly, caspase-8 is recruited to
the AIM2 inflammasome to drive apoptosis in Francisella-infected
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS 271
Table 1. Bacteria recognized by the AIM2 inflammasome
Bacteria Cells Mice
Francisella tularensis LVS
or F. tularensis subspecies
novicida (F. novicida)
r
Reduced caspase-1 activation, IL-1β
and IL-18, and pyroptosis in Aim2
/
mouse BMDMs or BMDCs
[26–28, 44–47].
r
Reduced IL-1β in shRNA-silenced
Aim2 THP1 human monocytic cell line
[48].
r
Increased overall susceptibility,
reduced serum IL-18 levels 1d p.i. and
increased bacterial burden 3d p.i.
[27, 44, 45].
Listeria monocytogenes
r
Reduced caspase-1 cleavage, IL-1β
release, and pyroptosis in
Aim2-silenced mouse BMDMs [49–52].
r
N/A
Streptococcus pneumoniae
r
Reduced caspase-1 cleavage, IL-1β
and IL-18 release, and pyroptosis in
Aim2-silenced peritoneal mouse
macrophages [53].
r
N/A
Mycobacterium tuberculosis
r
Reduced IL-1β and IL-18 release in
Aim2
/
and Aim2-silenced THP-1
human monocytic cell line [54, 55].
r
Increased overall susceptibility,
increased bacterial burden in the
lungs and liver 4 weeks p.i. [56].
r
Reduced caspase-1 cleavage and
increased infiltration of inflammatory
cells in lungs [56].
r
Reduced IL-1β in BALF and IL-18 in
serum at 3 weeks p.i. [56].
r
Reduced IFN-γ production by CD4
+
T cells [56].
Mycobacterium bovis
r
Reduced caspase-1 cleavage, IL-1β
release, and pyroptosis in Aim2
-silenced mouse macrophage J774A.1
[57].
r
N/A
Porphyromonas gingivalis
r
Reduced caspase-1 cleavage, IL-1β,
and pyroptosis in siRNA-silenced
Aim2 THP-1 human monocytic cell
line [58].
r
N/A
Legionella pneumophila
SdhA
r
Reduced caspase-1 cleavage and IL-1β
release in Aim2
/
mouse BMDMs [78].
r
N/A
Staphylococcus aureus
r
N/A
r
Increased susceptibility upon
intracranial infection [59].
r
Reduced IL-1β, IL-6, CCL2, and CXCL10
in wound abscess [59].
Brucella abortus
r
Reduced caspase-1 cleavage, IL-1β,
and cell death in Aim2
/
mouse
BMDMs [60].
r
Increased bacterial burden 4 weeks p.i
[60].
Chlamydia muridarum
r
Reduced IL-1β and IL-18 in Aim2
/
mouse BMDMs [61].
r
N/A
BALF: bronchoalveolar lavage fluid; BMDCs: bone marrow-derived dendritic cells; BMDMs: bone marrow-derived macrophages; N/A: information
not available; p.i.: postinfection.
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
272 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
Figure 2. Regulation of the activation of the AIM2 inflammasome. The AIM2 inflammasome is activated by a number of microbial pathogens
and dsDNA ligands, including the DNA virus MCMV, the cytosolic bacterium F. novicida, and the dsDNA ligand poly(dA:dT). MCMV infection or
transfection of poly(dA:dT) leads to “canonical” activation of the AIM2 inflammasome, which does not require the type I IFN pathway. F. novicida
infection activates the AIM2 inflammasome via a “noncanonical” pathway owing to its requirement for type I IFN, analogous to the non-canonical
NLRP3 inflammasome pathway. Intracellular F. novicida releases DNA into the cytoplasm to activate the DNA sensors cGAS, STING, and IFI204,
which drive transcription of genes encoding type I IFN molecules. It remains unclear why the released DNA is unable to activate AIM2 at this stage,
since AIM2 is constitutively expressed in the cell. Type I IFN provides a feedback loop to induce expression of the transcription factor IRF1, which
upregulates expression of the IFN-inducible GTPases, including GBP2 and GBP5. GBP2 and GBP5 are recruited to bacterial structures, however,
whether they directly target the bacterial membrane or the membrane of intact Francisella-containing vacuole is unclear. Nevertheless, GBPs
mediate bacterial killing, resulting in abundant release of bacterial DNA for recognition by AIM2. Assembly of the AIM2 inflammasome induces
caspase-1-dependent cleavage of pro-IL-1β and pro-IL-18. Caspase-1 also drives cleavage of the substrate gasdermin D to induce pyroptosis.
cells in the absence of caspase-1 [47, 62], suggesting a complex
interplay between members of the caspase family.
Activation of the AIM2 inflammasome by F. novicida and
F. tularensis LVS requires the ability of the bacteria to escape the
vacuole into the host cytoplasm, a process mediated by a range
of bacterial virulence factors, including the transcriptional regu-
lator MglA and proteins encoded by the Francisella-pathogenicity
island [26–28, 63, 64]. The Francisella-pathogenicity island is a
genomic region that contains a cluster of 16–19 genes encoding
virulence factors of the bacterium [65]. Similarly, L. monocyto-
genes must escape the vacuole and undergo bacteriolysis in order
to induce the activation of the AIM2 inflammasome [49–52, 66].
Type I IFN potentiates the activity of the AIM2 inflammasome dur-
ing bacterial infection [26–28, 67, 68]. In response to F. novicida
infection, the DNA sensors cGAS, IFI204, and STING cooperate
to detect small amounts of DNA released by the bacteria to drive
production of type I IFN in mouse macrophages [27, 46, 69]. Type
I IFN is then released to the outside of the cell, where it binds to
the type I IFN receptor (IFNR) in an autocrine manner, activat-
ing the IFN-stimulated gene factor 3 [70]. IFN-stimulated gene
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS 273
factor 3 is comprised of the transcription factors STAT1, STAT2,
and IRF9, which drives the transcription of IFN-stimulated genes
(ISGs) [70]. A study has now demonstrated that, during infection
with F. novicida, signaling via type I IFN induces the expression
of the transcription factor IRF1, where IRF1 further drives expres-
sion of IFN-inducible GTPases called guanylate-binding proteins
(GBPs) [27]. Induction of both IRF1 and GBPs are necessary to
fully engage the AIM2 inflammasome by F. novicida infection
(Fig. 2) [27, 28].
GBPs are clustered over two locations in the genome of mice.
Genes encoding GBP1, GBP2, GBP3, GBP5, and GBP7 are located
on chromosome 3, whereas genes encoding GBP4, GBP6, GBP8,
GBP9, GBP10, and GBP11 are found on chromosome 5 [71]. Of
these, GBP2 and GBP5 are recruited to cytoplasmic F. novicida
bacteria to drive bacterial killing, exposing abundant amounts of
bacterial DNA for detection by AIM2 [27, 28]. GBP2 and GBP5
function in a nonredundant manner, and reconstitution of either
GBP in type I IFN receptor 1-deficient macrophages cannot rescue
inflammasome activation, indicating that type I IFN signaling acti-
vates GBPs, possibly via expression other IFN-inducible proteins
[27, 28]. Overall, mice lacking AIM2, caspase-1, IRF1, or GBPs
have been shown to secrete reduced levels of IL-18 in response
to F. novicida infection and are all hypersusceptible to F. novicida
infection compared with wild-type mice [27, 28, 44–46]. In agree-
ment with this observation, antibody-mediated neutralization of
IL-1β and IL-18 in WT mice increases susceptibility to F. novicida
infection [64].
The precise mechanism of bacterial killing mediated by GBPs
remains unknown. Recent studies have shown that the antimicro-
bial activity of the gp91 subunit of NADPH oxidase (also known
as NOX2) and inducible nitric oxide synthase are not required for
mediating activation of the AIM2 inflammasome [28, 72]. How-
ever, the pharmacological inhibition of reactive oxygen species
(ROS) and mitochondrial ROS partially reduces caspase-1 acti-
vation, and therefore, the release of IL-1β driven by F. novicida
infection [28, 72]. ROS inhibition impairs the expression of IL-1β
and TNF, arguing that further evidence is required to convinc-
ingly link ROS-mediated bacterial killing and activation of the
AIM2 inflammasome [73]. It also remains a mystery as to why
DNA molecules that are released to activate cGAS, STING, and
IFI204 are unable to activate AIM2 at this stage. One possibility is
that the concentration of DNA that is sufficient to activate cGAS,
STING, and IFI204 is lower than the concentration required to
activate AIM2. AIM2 is constitutively expressed in the cell, but its
expression is also induced by IFN [11, 37, 38]. However, trans-
fection of dsDNA into the cytoplasm can directly activate AIM2
independently of IFN [27, 28, 73], ruling out a requirement for
“priming” in activation of the AIM2 inflammasome.
Bacteria have evolved virulence determinants to prevent
release of DNA and other bacterial ligands and avoid cytoplasmic
detection and clearance by inflammasomes. Francisella tularensis
subspecies tularensis SchuS4 have been shown to induce low levels
of inflammasome activation and IL-1β secretion in primary mouse
bone marrow-derived macrophages, possibly due to enhanced
resistance to H
2
O
2
to protect itself from bacteriolysis, or other
virulence factors that confer evasion of the immune system [72].
The putative lipid II flippase, MviN, and RipA, a protein used for
intracellular replication, of F. tularensis LVS are both required to
dampen AIM2 inflammasome responses [74, 75]. Further studies
have shown that F. tularensis LVS or F. novicida mutants lack-
ing MviN, RipA, and several membrane-associated proteins or
proteins involved in O-antigen or LPS biosynthesis are hypersus-
ceptible to intracellular lysis and DNA release in macrophages,
providing a rationale for why these mutants induce elevated
activation of the AIM2 inflammasome [63]. Genes encoding the
5-formyltetrahydrofolate cycloligase within the folate metabolic
pathway and pseudouridine synthase in F. tularensis LVS have also
been demonstrated to influence the magnitude of AIM2 inflam-
masome activation [76]. A more recent study identified a clus-
tered, regularly interspaced, short palindromic repeats-CRISPR
associated (CRISPR-Cas) system used by F. novicida to strengthen
the integrity of its bacterial membrane, leading to reduced DNA
release in the cytoplasm [77]. Another example is found in
Legionella pneumophila, which encodes an effector protein SdhA,
shown to prevent rupture of the Legionella-containing vacuole and
thereby minimizing the amount of bacterial DNA released into the
cytoplasm [78].
There is limited evidence so far to support the existence of
mechanisms used by bacteria to directly inhibit or evade activation
of the AIM2 inflammasome. However, F. tularensis LVS and the
virulent SchuS4 strain do suppress TLR2-dependent responses to
reduce the level of pro-IL-1β available for cleavage by the inflam-
masome [79]. Overall, the AIM2 inflammasome is an effective
antimicrobial machinery against certain bacterial pathogens.
Viral infection
Inflammasome responses play an essential role in the host pro-
tection against viral infection [80, 81]. Genetic materials from
DNA viruses that enter the cytoplasm can be detected by AIM2,
for instance mouse cytomegalovirus (MCMV), vaccinia virus, and
human papillomaviruses (Table 2) [11, 44, 82]. MCMV and vac-
cinia viruses robustly induce inflammasome responses in mouse
macrophages in an AIM2-dependent manner (Fig. 2) [11, 27, 44].
Further, Aim2
/
mice infected with MCMV have an impaired abil-
ity to secrete IL-18, carry a higher viral titre, and have reduced lev-
els of IFN-γ-producing NK cells compared with infected WT mice
[44]. Human papillomaviruses have also been shown to drive IL-1β
and IL-18 release in human keratinocytes in an AIM2-dependent
manner [82].
To date, there is no strong evidence in the literature to indi-
cate that DNA viruses other than MCMV and vaccinia virus
activate the AIM2 inflammasome. A recent study suggests that
in the human glomerular mesangial cell line infected with
hepatitis B virus, siRNA-mediated silencing of the gene encoding
AIM2 leads to a reduced expression of IL-1β, IL-18, and caspase-
1 [83]. Whether AIM2 directly mediates the recognition of viral
DNA derived from hepatitis B virus in immune or nonimmune cells
has not been established. Comparison of the expression of AIM2 in
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
274 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
Table 2. Viruses recognized by the AIM2 inflammasome
Viruses Cells Mice
MCMV (DNA virus)
r
Reduced caspase-1 cleavage
and IL-1β release in Aim2
/
mouse peritoneal
macrophages BMDMs and
BMDCs [27, 44].
r
Increased viral titers 36 h p.i. [44].
r
Reduced serum IL-18 level 36 h p.i. [44].
r
Reduced IFN-γ production [44].
Vaccinia virus (DNA virus)
r
Reduced caspase-1 cleavage
and IL-1β release in Aim2
/
mouse peritoneal
macrophages and BMDCs [44].
r
N/A
Human papillomaviruses (DNA virus)
r
Reduced IL-1β and IL-18
release in Aim2-silenced
human keratinocytes [82].
r
N/A
Hepatitis B virus (DNA virus)
r
Reduced gene expression of
IL-1β, IL-18, and caspase-1 in
Aim2-silenced human
glomerular mesangial cell line
[83].
r
N/A
Chikungunya virus (RNA virus)
r
Reduced IL-1β release in
Aim2-silenced human primary
dermal fibroblasts [92].
r
N/A
West Nile virus (RNA virus)
r
Reduced IL-1β release in
Aim2-silenced primary human
dermal fibroblasts [92].
r
N/A
BMDCs: bone marrow-derived dendritic cells; BMDMs: bone marrow-derived macrophages; N/A: information not available; p.i.: postinfection.
patients with acute and chronic hepatitis B revealed that those at
the acute stage expressed higher levels of AIM2 in peripheral blood
mononuclear cells [84]. A subsequent study reported that 89.4%
of the liver tissues collected from individuals with chronic hepatitis
B virus infection were positive for AIM2 expression by immuno-
histochemistry, compared with only 8.7% of those with chronic
hepatitis C infection [85]. Expression of the gene encoding AIM2
has been reported to be significantly higher in kidney tissues of
patients with hepatitis B virus-associated glomerulonephritis com-
pared with patients with chronic glomerulonephritis [83].
In all cases, it is probable that viral DNA binds directly to
AIM2 to trigger inflammasome activation, but the precise molecu-
lar mechanism that leads to exposure of the viral DNA for sensing
by AIM2 is not entirely clear. Unlike F. novicida infection, type I
IFN signaling, IRF1, and GBPs are dispensable for the activation of
the AIM2 inflammasome by either MCMV infection or transfected
dsDNA [27]. However, AIM2 does not respond to all DNA viruses.
For example, adenovirus and herpesviruses HSV-1 and MHV-68
activate the NLRP3 or putative IFI16 inflammasome, rather than
the AIM2 inflammasome in mouse BM-derived macrophages or
mouse thioglycollate-elicited macrophages [8, 44, 80, 81]. During
HSV-1 infection of human macrophages, the capsid that encapsu-
lates the viral DNA is degraded by the proteasome, which releases
DNA for recognition by IFI16 in the cytoplasm [86], suggesting
that the inability of AIM2 to sense viral DNA is probably not due
to a lack of viral DNA release in the cytoplasm. It might be possi-
ble that certain DNA viruses can strategically inhibit the ability of
AIM2 to interact with their DNA. Alternatively, the DNA of certain
viruses, such as Hepatitis B virus and other Hepadnaviruses, could
be transcribed into RNA templates, which may serve as activators
for NLRP3 [87–89]. Indeed, a precedent exists for indirect sensing
of DNA by the RNA sensor RIG-I [90, 91]. RNA polymerase III tran-
scribes AT-rich DNA into dsRNA transcripts carrying an uncapped
5
-triphosphate moiety, which has been shown to activate RIG-I
[90, 91]. Conversely, a study has reported a role for AIM2 in driv-
ing IL-1β secretion in response to the RNA viruses [92]. Silencing
of genes encoding AIM2 and caspase-1 reduces proteolytic cleav-
age and release of IL-1β in human dermal fibroblasts infected with
the RNA viruses Chikungunya virus or West Nile virus [92]. How
AIM2 might sense RNA viruses is still unclear, and further charac-
terization of the molecular mechanism involved in the activation
of the AIM2 inflammasome by viruses is required.
Other pathogens
In addition to bacteria and viruses, AIM2 has been shown to
mediate pathogen recognition of, and host defense to, the fun-
gal pathogen Aspergillus fumigatus and the protozoan Plasmodium
berghei (Table 3) [93, 94]. AIM2 and NLRP3 function in a redun-
dant fashion to confer inflammasome activation against A. fumi-
gatus infection in mouse BM-derived DCs and mice [93] Further-
more, mice lacking both AIM2 and NLRP3, ASC or caspase-1, and
infected with A. fumigatus, are more susceptible than infected
WT mice [93]. The requirement for dual sensing of pathogens
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS 275
Table 3. Fungi and parasites recognized by the AIM2 inflammasome
Cells Mice
Fungi
Aspergillus fumigatus Slight reduction in IL-1β and IL-18
in Aim2
/
mouse BMDCs owing
to redundant roles with NLRP3
[93].
Not susceptible owing to
redundant roles with NLRP3 [93].
Protozoa
Plasmodium berghei Slight reduction in IL-1β and
pyroptosis in Aim2
/
mouse
BMDMs infected with iRBCs,
synthetic and natural hemozoin
owing to redundant roles with
NLRP3 [94].
Decreased neutrophils recruitment
in peritoneal cavity owing to
redundant roles with NLRP3,
after 15 h p.i. [94].
BMDCs: bone marrow-derived dendritic cells; iRBCs: infected red blood cells.
by both AIM2 and NLRP3 has also been observed in mouse
BM-derived macrophages stimulated with Plasmodium berghei-
infected red blood cells or synthetic and natural hemozoins [94].
AIM2 also directly recognizes A. fumigatus genomic DNA intro-
duced into the cytoplasm by a transfection agent [93] or P. falci-
parum genomic DNA transported into the cytoplasm by hemozoins
[94]. It would be interesting to identify additional pathogens that
can activate the AIM2 inflammasome.
AIM2 role in cancer biology
AIM2 has also been shown to suppress the development of can-
cer [95, 96]. The gene encoding AIM2 was originally isolated
from human melanoma cells [97]. Reduced expression and fre-
quent frameshift microsatellite instability of the AIM2 have been
observed in tumor tissues from patients with colorectal cancer
[98–101]. Colorectal cancer patients whose tissues have reduced
AIM2 expression have a poorer prognosis compared with those
with a higher level of AIM2 expression [98]. Reduced expression
of AIM2 has also been reported in prostate cancer [102], whereas
increased expression has been detected in nasopharyngeal carci-
noma tumors [103, 104], oral squamous cell carcinoma [105],
and lung adenocarcinoma [106]. The differential expression of
AIM2 in a range of tumor tissues suggests that it may have unique
roles in different types of cancer.
The mechanism for AIM2 in the regulation of tumorigenesis
has been described in a mouse model of colitis-associated col-
orectal cancer [95, 96]. Two groups have recently demonstrated
that AIM2 operates independently of the inflammasome to prevent
colorectal cancer [95, 96]. Both studies found that Aim2
/
mice
developed severe colitis, polyps, and higher tumor burden upon
administration of AOM and DSS [95, 96]. Although differential
production of the major inflammatory mediators, including TNF
and IL-6, was not observed between WT and Aim2
/
mice, pro-
liferation of enterocytes was more pronounced in Aim2
/
mice
[95, 96]. Indeed, murine fibroblasts and colon cancer cell lines
expressing an AIM2-encoding construct have been shown to have
an impaired ability to undergo proliferation [40, 107]. Overex-
pression of AIM2 in colon cancer cell lines also induces cell cycle
arrest, and the transformed cells exhibit a delayed transition from
the late S-phase to the G2/M phase [107], suggesting that AIM2
plays a proliferation-inhibitory role in these cancer cell lines.
Furthermore, in another mouse model of intestinal cancer,
Aim2
/
mice carrying aberrant activating β-catenin mutations
failed to prevent the expansion of cancer-associated stem cells
in the small and large intestine [95]. Similarly, Aim2
/
mice
harboring the heterozygous mutation in the adenomatous poly-
posis coli gene developed more tumors than Aim2
+/+
mice car-
rying the mutant adenomatous polyposis coli gene [96]. Intesti-
nal stem cells lacking AIM2 proliferated more than WT intestinal
stem cells in organoid culture, and this proliferation was associ-
ated with increased activation of the kinase AKT [95, 96]. Wilson
and colleagues [108, 109] identified DNA-PK, a kinase that can
phosphorylate and activate AKT, as a binding partner of AIM2,
whereby AIM2 suppresses the activation of DNA-PK- and DNA-PK-
dependent phosphorylation of AKT at the serine residue 473 [96].
Indeed, treatment of Aim2
/
mice with an AKT inhibitor reduced
tumor burden in Aim2
/
mice with colitis [96].
Our laboratory performed further analysis and found that
Aim2
/
mice harbored a microbial ecology different to that of
WT mice [95]. Reciprocal exchange of the microbiota between
Aim2
/
and WT mice by means of cohousing substantially
reduced tumorigenesis in Aim2
/
mice and increased tumorige-
nesis in WT mice [95]. Collectively, these studies provide insights
into the function of AIM2 in colorectal cancer, and highlight that
potential therapies that inhibit the AKT pathway can be further
investigated for treatment of cancer associated with AIM2 muta-
tions [40, 95, 96, 107].
AIM2 in inflammatory, autoimmune,
and other pathological conditions
Given that the host DNA is normally sequestered in the nucleus or
mitochondria, the accumulation of host DNA in the cytosol, due to
impaired degradation or clearance or excess uptake of extracellu-
lar DNA from dying neighboring cells, could induce inflammation.
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
276 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
For instance, accumulated DNA can serve as an endogenous dan-
ger signal, and has been shown to trigger AIM2-dependent release
of IL-1β in skin cells, contributing to the pathogenesis of psoriasis
[38]. Scavenging of DNA by the antimicrobial cathelicidin pep-
tide LL-37 produced by the inflamed skin of psoriasis patients
prevents overt activation of the AIM2 inflammasome and IL-1β
release [38]. Increased AIM2 expression has been observed in
patients with acute and chronic skin conditions, including pso-
riasis, atopic dermatitis, venous ulcera, contact dermatitis, and
experimental wounds in humans [38, 110]. In addition, expres-
sion of the gene encoding AIM2 is elevated in immune cells of
male patients with systemic lupus erythematosus (SLE) and both
increases and decreases in AIM2 expression have been observed
in female patients with SLE [111, 112]. Further, DNA methylation
of the gene encoding AIM2 is reduced in patients with SLE com-
pared with their healthy siblings [113], suggesting that differential
expression or epigenetic changes could be linked to development
of the disease.
Increased expression of AIM2 has been reported in patients
with inflammatory bowel diseases and liver inflammation [114–
116]. For example, elevated expression of AIM2 has been detected
in ascitic fluid macrophages collected from cirrhotic patients,
compared with PBMCs from the same patients, or with CD14
+
macrophages from peripheral blood mononuclear cells of healthy
individuals [115]. Increased expression of AIM2 and NLRP3 and
elevated activation of caspase-1 and maturation of IL-1β have been
found in liver tissues of mice with steatohepatitis [116].
Moreover, there is evidence to suggest that AIM2 is involved in
inflammation and cell death of the brain. Cell-free DNA fragments
are more frequently detected in the cerebrospinal fluid (CSF) of
patients with traumatic brain injury than in CSF from nontrauma
patients [117]. When human embryonic cortical neurons, which
express AIM2 inflammasome components and have been shown
to be capable of IL-1β release and cell death upon AIM2 activa-
tion with poly(dA:dT) transfection [117], were exposed to the
CSF of traumatic brain injury patients, they exhibited increased
AIM2 expression and caspase-1 activation compared with embry-
onic cortical neurons that had been exposed to the CSF of non-
trauma patients [117]. A further study has shown that mice lacking
AIM2 are more protected than WT mice to ischemic brain injury
[118]. Upon induction of focal cerebral ischemia, less activation
of microglial cells and recruitment of leukocytes were found in
Aim2
/
mice compared with WT mice [118].
The inability to degrade self-DNA also contributes to the patho-
genesis of autoimmune polyarthritis. Mice lacking the lysoso-
mal endonuclease DNAse II (Dnase II
/
mice) are embryoni-
cally lethal, owing to an impaired ability to degrade self-DNA
by macrophages [119]. Genomic deletion of type I IFN recep-
tor (IFNAR) rescued Dnase II
/
mice from embryonic lethal-
ity [120]; however, the mice lacking both IFNAR and DNAse II
(Dnase II
/
Ifnar
/
mice) would eventually develop polyarthri-
tis [121]. Intriguingly, genomic deletion of AIM2 was shown to
prevent inflammasome activation, inflammatory cytokine produc-
tion, macrophage infiltration in the joint, and the development of
arthritis in Dnase II
/
Ifnar
/
mice [122, 123]. Genomic dele-
tion of STING was also shown to protect Dnase II
/
Ifnar
/
mice
from joint inflammation [122], indicating that multiple DNA sen-
sors might contribute to the inappropriate DNA recognition driv-
ing clinical manifestation. Overall, aberrant activation of AIM2
from self-DNA is a key driver of inflammatory and autoimmune
diseases.
Conclusions
A wide range of microbial pathogens is sensed by AIM2 in mam-
malian cells. Recognition of DNA from pathogens by AIM2 leads to
protective inflammasome-mediated host responses. Recent stud-
ies have demonstrated that AIM2 inflammasome also plays impor-
tant roles in nonmicrobial diseases, highlighting the multifaceted
nature of AIM2 beyond immunity to infectious diseases. In colorec-
tal cancer, AIM2 orchestrates inflammasome-independent func-
tions by suppressing stem cell proliferation, and contributes to
maintenance of a healthy gut microbiota. The role of AIM2 inflam-
masome in other types of cancer should be further explored.
Moreover, inappropriate recognition of self-DNA by AIM2 trig-
gers detrimental inflammatory responses, leading to superficial
and systemic inflammation. Inhibiting AIM2 inflammasome activ-
ity using synthetic inhibitors, such as suppressive oligodeoxynu-
cleotides [124], or harnessing the power of endogenous AIM2
inhibitors, such as pyrin-containing proteins [36, 39] or antimi-
crobial cathelicidin peptides [38], could be investigated for their
potential to control unwanted inflammation. An additional line of
research could focus on understanding the complementary rela-
tionship between AIM2 and a plethora of other DNA sensors in
the context of different cell types, tissues, and organs. A holis-
tic understanding of the biology of AIM2 could lead to improved
immunosurveillance in the fight against infectious diseases and
cancer, while avoiding debilitating inflammatory and autoimmune
diseases.
Acknowledgments: Research studies from the lab are supported
by the US National Institutes of Health (AR056296, CA163507,
and AI101935 to T.-D.K.), the American Lebanese Syrian Asso-
ciated Charities (to T.-D.K.), and the R.G. Menzies Early Career
Fellowship from the National Health and Medical Research Coun-
cil of Australia (to S.M.M.).
Conflict of interest: The authors declare no financial or commer-
cial conflict of interest.
References
1 Hemmi, H., Takeuchi, O., Kawai, T., Kaisho, T., Sato, S., Sanjo, H., Mat-
sumoto, M. et al., A Toll-like receptor recognizes bacterial DNA. Nature
2000. 408: 740–745.
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS 277
2 Ishii, K. J., Coban, C., Kato, H., Takahashi, K., Torii, Y., Takeshita, F.,
Ludwig, H. et al., A Toll-like receptor-independent antiviral response
induced by double-stranded B-form DNA. Nat. Immunol. 2006. 7: 40–48.
3 Stetson, D. B. and Medzhitov, R., Recognition of cytosolic DNA activates
an IRF3-dependent innate immune response. Immunity 2006. 24: 93–103.
4 Atianand, M. K. and Fitzgerald, K. A., Molecular basis of DNA recogni-
tion in the immune system. J. Immunol. 2013. 190: 1911–1918.
5 Bhat, N. and Fitzgerald, K. A., Recognition of cytosolic DNA by cGAS and
other STING-dependent sensors. Eur. J. Immunol. 2014. 44: 634–640.
6 Cavlar, T., Ablasser, A. and Hornung, V., Induction of type I IFNs by
intracellular DNA-sensing pathways. Immunol. Cell Biol. 2012. 90: 474–
482.
7 Cai, X., Chiu, Y. H. and Chen, Z. J., The cGAS-cGAMP-STING pathway of
cytosolic DNA sensing and signaling. Mol. Cell 2014. 54: 289–296.
8 Muruve, D. A., Petrilli, V., Zaiss, A. K., White, L. R., Clark, S. A., Ross, P.
J., Parks, R. J. et al., The inflammasome recognizes cytosolic microbial
and host DNA and triggers an innate immune response. Nature 2008.
452: 103–107.
9 Stacey, K. J., Ross, I. L. and Hume, D. A., Electroporation and DNA-
dependent cell death in murine macrophages. Immunol. Cell Biol. 1993.
71 (Pt 2): 75–85.
10 Man, S. M. and Kanneganti, T. D., Regulation of inflammasome activa-
tion. Immunol. Rev.
2015. 265: 6–21.
11 Hornung, V., Ablasser, A., Charrel-Dennis, M., Bauernfeind, F., Horvath,
G., Caffrey,D.R., Latz, E. et al., AIM2 recognizes cytosolic dsDNA and
forms a caspase-1-activating inflammasome with ASC. Nature 2009. 458:
514–518.
12 Fernandes-Alnemri, T., Yu, J. W., Datta, P., Wu, J. and Alnemri, E. S.,
AIM2 activates the inflammasome and cell death in response to cyto-
plasmic DNA. Nature 2009. 458: 509–513.
13 Burckstummer, T., Baumann, C., Bluml, S., Dixit, E., Durnberger, G.,
Jahn, H., Planyavsky, M. et al., An orthogonal proteomic-genomic
screen identifies AIM2 as a cytoplasmic DNA sensor for the inflam-
masome. Nat. Immunol. 2009. 10: 266–272.
14 Roberts, T. L., Idris, A., Dunn, J. A., Kelly, G. M., Burnton, C. M., Hodgson,
S., Hardy, L. L. et al., HIN-200 proteins regulate caspase activation in
response to foreign cytoplasmic DNA. Science 2009. 323: 1057–1060.
15 Unterholzner, L., Keating, S. E., Baran, M., Horan, K. A., Jensen, S. B.,
Sharma, S., Sirois, C. M. et al., IFI16 is an innate immune sensor for
intracellular DNA. Nat. Immunol. 2010. 11: 997–1004.
16 Brunette, R. L., Young, J. M., Whitley, D. G., Brodsky, I. E., Malik, H.
S. and Stetson, D. B., Extensive evolutionary and functional diversity
among mammalian AIM2-like receptors. J. Exp. Med. 2012. 209: 1969–
1983.
17 Cridland, J. A., Curley, E. Z., Wykes, M. N., Schroder, K., Sweet, M. J.,
Roberts, T. L., Ragan, M. A. et al., The mammalian PYHIN gene family:
phylogeny, evolution and expression. BMC Evol. Biol. 2012. 12: 140.
18 Jin, T., Perry, A., Jiang, J., Smith, P., Curry, J. A., Unterholzner, L., Jiang,
Z. et al., Structures of the HIN domain: DNA complexes reveal ligand
binding and activation mechanisms of the AIM2 inflammasome and
IFI16 receptor. Immunity 2012. 36: 561–571.
19 Jin, T., Perry, A., Smith, P., Jiang, J. and Xiao, T. S., Structure of the
absent in melanoma 2 (AIM2) pyrin domain provides insights into the
mechanisms of AIM2 autoinhibition and inflammasome assembly. J.
Biol. Chem. 2013. 288: 13225–13235.
20 Lu, A., Kabaleeswaran, V., Fu, T., Magupalli, V. G. and Wu, H.,Crys-
tal structure of the F27G AIM2 PYD mutant and similarities of its self-
association to DED/DED interactions. J. Mol. Biol. 2014. 426: 1420–1427.
21 Hou, X. and Niu, X., The NMR solution structure of AIM2 PYD domain
from Mus musculus reveals a distinct alpha2-alpha3 helix conformation
from its human homologues. Biochem. Biophys. Res. Commun. 2015. 461:
396–400.
22 Lu, A., Magupalli, V. G.
, Ruan, J., Yin, Q., Atianand, M. K., Vos, M. R.,
Schroder, G. F. et al., Unified polymerization mechanism for the assem-
bly of ASC-dependent inflammasomes. Cell 2014. 156: 1193–1206.
23 Cai, X., Chen, J., Xu, H., Liu, S., Jiang, Q. X., Halfmann, R. and Chen, Z.
J., Prion-like polymerization underlies signal transduction in antiviral
immune defense and inflammasome activation. Cell 2014. 156: 1207–
1222.
24 Broz, P., Newton, K., Lamkanfi, M., Mariathasan, S., Dixit, V. M. and
Monack, D. M., Redundant roles for inflammasome receptors NLRP3
and NLRC4 in host defense against Salmonella. J. Exp. Med. 2010. 207:
1745–1755.
25 Man, S. M., Hopkins, L. J., Nugent, E., Cox, S., Gluck, I. M., Tourlomousis,
P., Wright, J. A. et al., Inflammasome activation causes dual recruitment
of NLRC4 and NLRP3 to the same macromolecular complex. Proc. Natl.
Acad. Sci. USA 2014. 111: 7403–7408.
26 Belhocine, K. and Monack, D. M., Francisella infection triggers activation
of the AIM2 inflammasome in murine dendritic cells. Cell Microbiol. 2012.
14: 71–80.
27 Man, S. M., Karki, R., Malireddi, R. K., Neale, G., Vogel, P., Yamamoto,
M.
, Lamkanfi, M. et al., The transcription factor IRF1 and guanylate-
binding proteins target activation of the AIM2 inflammasome by Fran-
cisella infection. Nat. Immunol. 2015. 16: 467–475.
28 Meunier, E., Wallet, P., Dreier, R. F., Costanzo, S., Anton, L., Ruhl, S.,
Dussurgey, S. et al., Guanylate-binding proteins promote activation of
the AIM2 inflammasome during infection with Francisella novicida. Nat.
Immunol. 2015. 16: 476–484.
29 Juruj, C., Lelogeais, V., Pierini, R., Perret, M., Py, B. F., Jamilloux, Y.,
Broz, P. et al., Caspase-1 activity affects AIM2 speck formation/stability
through a negative feedback loop. Front. Cell. Infect. Microbiol. 2013. 3: 14.
30 Morrone, S. R., Matyszewski, M., Yu, X., Delannoy, M., Egelman, E. H.
and Sohn, J., Assembly-driven activation of the AIM2 foreign-dsDNA
sensor provides a polymerization template for downstream ASC. Nat.
Commun. 2015. 6: 7827.
31 Shi, J., Zhao, Y., Wang, K., Shi, X., Wang, Y., Huang, H., Zhuang, Y. et al.,
Cleavage of GSDMD by inflammatory caspases determines pyroptotic
cell death. Nature 2015. 526: 660–665.
32 Kayagaki, N., Stowe, I. B., Lee, B. L., O’Rourke, K.,
Anderson, K., Warm-
ing, S., Cuellar, T. et al., Caspase-11 cleaves gasdermin D for non-
canonical inflammasome signaling. Nature 2015. 526: 666–671.
33 Hara, H., Tsuchiya, K., Kawamura, I., Fang, R., Hernandez-Cuellar, E.,
Shen, Y., Mizuguchi, J. et al., Phosphorylation of the adaptor ASC acts as
a molecular switch that controls the formation of speck-like aggregates
and inflammasome activity. Nat. Immunol. 2013. 14: 1247–1255.
34 Rodgers, M. A., Bowman, J. W., Fujita, H., Orazio, N., Shi, M., Liang,
Q., Amatya, R. et al., The linear ubiquitin assembly complex (LUBAC)
is essential for NLRP3 inflammasome activation. J. Exp. Med. 2014. 211:
1333–1347.
35 Shi, C. S., Shenderov, K., Huang, N. N., Kabat, J., Abu-Asab, M., Fitzger-
ald,K.A., Sher, A. et al., Activation of autophagy by inflammatory sig-
nals limits IL-1beta production by targeting ubiquitinated inflamma-
somes for destruction. Nat. Immunol. 2012. 13: 255–263.
36 Khare, S., Ratsimandresy, R. A., de Almeida, L., Cuda, C. M., Rellick,
S. L., Misharin, A. V., Wallin, M. C. et al., The PYRIN domain-only protein
POP3 inhibits ALR inflammasomes and regulates responses to infection
with DNA viruses. Nat. Immunol. 2014. 15: 343–353.
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
278 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
37 Veeranki, S., Duan, X., Panchanathan, R. , Liu, H. and Choubey, D., IFI16
protein mediates the anti-inflammatory actions of the type-I interferons
through suppression of activation of caspase-1 by inflammasomes. PLoS
One 2011. 6: e27040.
38 Dombrowski, Y., Peric, M., Koglin, S., Kammerbauer, C., Goss, C., Anz,
D., Simanski, M. et al., Cytosolic DNA triggers inflammasome activation
in keratinocytes in psoriatic lesions. Sci. Transl. Med. 2011. 3: 82ra38.
39 de Almeida, L., Khare, S., Misharin, A. V., Patel, R., Ratsimandresy, R.
A., Wallin, M. C., Perlman, H. et al., T he PYRIN domain-only protein
POP1 inhibits inflammasome assembly and ameliorates inflammatory
disease. Immunity 2015. 43: 264–276.
40 Choubey, D., Walter, S., Geng, Y. and Xin, H., Cytoplasmic localization
of the interferon-inducible protein that is encoded by the AIM2 (absent
in melanoma) gene from the 200-gene family. FEBS Lett. 2000. 474:
38–42.
41 Yin, Q., Sester, D. P., Tian, Y., Hsiao, Y. S., Lu, A., Cridland, J. A., Sagu-
lenko, V. et al., Molecular mechanism for p202-mediated specific inhi-
bition of AIM2 inflammasome activation. Cell Rep. 2013. 4: 327–339.
42 Choubey, D. and Panchanathan, R., Interferon-inducible Ifi200-family
genes in systemic lupus erythematosus. Immunol. Lett. 2008. 119: 32–41.
43 Sester, D. P.
, Sagulenko, V., Thygesen, S. J., Cridland, J. A., Loi, Y. S.,
Cridland, S. O., Masters, S. L. et al., Deficient NLRP3 and AIM2 inflam-
masome function in autoimmune NZB mice. J. Immunol. 2015. 195: 1233–
1241.
44 Rathinam, V. A., Jiang, Z., Waggoner, S. N., Sharma, S., Cole,L.E., Wag-
goner, L., Vanaja,S.K.etal., The AIM2 inflammasome is essential for
host defense against cytosolic bacteria and DNA viruses. Nat. Immunol.
2010. 11: 395–402.
45 Fernandes-Alnemri, T., Yu,J.W., Juliana, C., Solorzano, L., Kang, S.,
Wu, J., Datta, P. et al., The AIM2 inflammasome is critical for innate
immunity to Francisella tularensis. Nat. Immunol. 2010. 11: 385–393.
46 Jones, J. W., Kayagaki, N., Broz, P., Henry, T., Newton, K., O’Rourke, K.,
Chan, S. et al., Absent in melanoma 2 is required for innate immune
recognition of Francisella tularensis. Proc. Natl. Acad. Sci. USA 2010. 107:
9771–9776.
47 Pierini, R., Juruj, C., Perret, M., Jones, C. L., Mangeot, P., Weiss, D. S.
and Henry, T., AIM2/ASC triggers caspase-8-dependent apoptosis in
Francisella-infected caspase-1-deficient m acrophages.
Cell Death Differ.
2012. 19: 1709–1721.
48 Atianand, M. K., Duffy,E.B., Shah, A., Kar, S., Malik, M. and Harton, J.
A., Francisella tularensis reveals a disparity between human and mouse
NLRP3 inflammasome activation. J. Biol. Chem. 2011. 286: 39033–39042.
49 Kim, S., Bauernfeind, F., Ablasser, A., Hartmann, G., Fitzgerald, K. A.,
Latz, E. and Hornung, V., Listeria monocytogenes is sensed by the NLRP3
and AIM2 inflammasome. Eur. J. Immunol. 2010. 40: 1545–1551.
50 Sauer, J. D., Witte, C. E., Zemansky, J ., Hanson, B., Lauer, P. and Portnoy,
D. A., Listeria monocytogenes triggers AIM2-mediated pyroptosis upon
infrequent bacteriolysis in the macrophage cytosol. Cell Host Microbe
2010. 7: 412–419.
51 Warren, S. E., Armstrong, A., Hamilton, M. K., Mao, D. P., Leaf, I. A.,
Miao, E. A. and Aderem, A., Cutting edge: cytosolic bacterial DNA acti-
vates the inflammasome via Aim2. J. Immunol. 2010. 185: 818–821.
52 Wu, J., Fernandes-Alnemri, T. and Alnemri, E. S., Involvement of the
AIM2, NLRC4, and NLRP3 inflammasomes in caspase-1 activation by
Listeria monocytogenes. J. Clin. Immunol. 2010. 30: 693–702.
53 Fang, R., Tsuchiya, K., Kawamura, I., Shen, Y., Hara, H.
, Sakai, S.,
Yamamoto, T. et al., Critical roles of ASC inflammasomes in caspase-1
activation and host innate resistance to Streptococcus pneumoniae infec-
tion. J. Immunol. 2011. 187: 4890–4899.
54 Wassermann, R., Gulen, M. F., Sala, C., Perin, S. G., Lou, Y., Rybniker,
J., Schmid-Burgk, J. L. et al., Mycobacterium tuberculosis differentially
activates cGAS- and inflammasome-dependent intracellular immune
responses through ESX-1. Cell Host Microbe 2015. 17: 799–810.
55 Saiga, H., Nieuwenhuizen, N., Gengenbacher, M., Koehler, A. B.,
Schuerer, S., Moura-Alves, P., Wagner, I. et al., The recombinant
BCG ureC::hly vaccine targets the AIM2 inflammasome to induce
autophagy and inflammation. J. Infect. Dis. 2015. 211: 1831–1841.
56 Saiga, H., Kitada, S., Shimada, Y., Kamiyama, N., Okuyama, M., Makino,
M., Yamamoto, M. et al., Critical role of AIM2 in Mycobacterium tubercu-
losis infection. Int. Immunol. 2012. 24: 637–644.
57 Yang, Y., Zhou, X., Kouadir, M., Shi, F., Ding, T., Liu, C., Liu, J. et al.,
The AIM2 inflammasome is involved in macrophage activation during
infection with virulent Mycobacterium bovis strain. J. Infect. Dis.
2013. 208:
1849–1858.
58 Park, E., Na, H. S., Song,Y.R., Shin, S. Y., Kim, Y. M. and Chung, J.,Acti-
vation of NLRP3 and AIM2 inflammasomes by Porphyromonas gingivalis
infection. Infect. Immun. 2014. 82: 112–123.
59 Hanamsagar, R., Aldrich, A. and Kielian, T., Critical role for the AIM2
inflammasome during acute CNS bacterial infection. J. Neurochem. 2014.
129: 704–711.
60 Gomes, M. T., Campos,P.C., Oliveira, F. S., Corsetti,P.P., Bortoluci, K.
R., Cunha, L. D., Zamboni, D. S. et al., Critical role of ASC inflamma-
somes and bacterial type IV secretion system in caspase-1 activation
and host innate resistance to Brucella abortus infection. J. Immunol. 2013.
190: 3629–3638.
61 Finethy, R., Jorgensen, I., Haldar, A. K., de Zoete, M. R., Strowig, T.,
Flavell, R. A., Yamamoto, M. et al., Guanylate binding proteins enable
rapid activation of canonical and noncanonical inflammasomes in
chlamydia-infected macrophages. Infect. Immun. 2015. 83: 4740–4749.
62 Sagulenko, V., Thygesen, S. J., Sester, D. P., Idris, A., Cridland, J. A.,
Vajjhala, P. R., Roberts, T. L. et al., AIM2 and NLRP3 inflammasomes
activate both apoptotic and pyroptotic death pathways via ASC. Cell
Death Differ. 2013. 20: 1149–1160.
63 Peng, K., Broz, P.,
Jones, J., Joubert, L. M. and Monack, D., Elevated AIM2-
mediated pyroptosis triggered by hypercytotoxic Francisella mutant
strains is attributed to increased intracellular bacteriolysis. Cell Micro-
biol. 2011. 13: 1586–1600.
64 Mariathasan, S., Weiss, D. S., Dixit, V. M. and Monack, D. M.,
Innate immunity against Francisella tularensis is dependent on the
ASC/caspase-1 axis. J. Exp. Med. 2005. 202: 1043–1049.
65 Nano, F. E. and Schmerk, C.,TheFrancisella pathogenicity island. Ann.
NY Acad. Sci. 2007. 1105: 122–137.
66 Tsuchiya, K., Hara, H., Kawamura, I., Nomura, T., Yamamoto, T., Daim,
S., Dewamitta, S. R. et al., Involvement of absent in melanoma 2 in
inflammasome activation in macrophages infected with Listeria mono-
cytogenes. J. Immunol. 2010. 185: 1186–1195.
67 Henry, T., Brotcke, A., Weiss, D. S., Thompson, L. J. and Monack, D. M.,
Type I interferon signaling is required for activation of the inflamma-
some during Francisella infection. J. Exp. Med. 2007. 204: 987–994.
68 Fang, R., Hara, H., Sakai, S., Hernandez-Cuellar, E., Mitsuyama, M.,
Kawamura, I. and Tsuchiya, K., Type I interferon signaling regulates
activation of the absent in melanoma 2 inflammasome during Strepto-
coccus pneumoniae infection. Infect. Immun. 2014. 82: 2310–2317.
69 Storek, K. M.
, Gertsvolf,N.A., Ohlson, M. B. and Monack, D. M.,cGAS
and Ifi204 cooperate to produce type I IFNs in response to Francisella
infection. J. Immunol. 2015. 194: 3236–3245.
70 Ivashkiv, L. B. and Donlin, L. T., Regulation of type I interferon
responses. Nat. Rev. Immunol. 2014. 14: 36–49.
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
Eur. J. Immunol. 2016. 46: 269–280 HIGHLIGHTS 279
71 Yamamoto, M., Okuyama, M., Ma, J. S., Kimura, T., Kamiyama, N.,
Saiga, H., Ohshima, J. et al., A cluster of interferon-gamma-inducible
p65 GTPases plays a critical role in host defense against Toxoplasma
gondii. Immunity 2012. 37: 302–313.
72 Crane, D. D., Bauler, T. J., Wehrly, T. D. and Bosio, C. M., Mitochondrial
ROS potentiates indirect activation of the AIM2 inflammasome. Front.
Microbiol. 2014. 5: 438.
73 Bauernfeind, F., Bartok, E., Rieger, A., Franchi, L., Nunez, G. and Hor-
nung, V., Cutting edge: reactive oxygen species inhibitors block priming,
but not activation, of the NLRP3 inflammasome. J. Immunol. 2011. 187:
613–617.
74 Huang, M. T., Mortensen, B. L., Taxman, D. J., Craven, R. R., Taft-Benz,
S., Kijek, T. M., Fuller, J. R. et al., Deletion of ripA alleviates suppression
of the inflammasome and MAPK by Francisella tularensis. J. Immunol. 2010.
185: 5476–5485.
75 Ulland, T. K., Buchan, B. W., Ketterer, M. R., Fernandes-Alnemri, T.,
Meyerholz, D. K., Apicella, M. A., Alnemri, E. S. et al., Cutting edge:
mutation of Francisella tularensis mviN leads to increased macrophage
absent in melanoma 2 inflammasome activation and a loss of virulence.
J. Immunol. 2010. 185: 2670–2674.
76 Ulland, T. K.
, Janowski, A. M., Buchan, B. W., Faron, M., Cassel, S.
L., Jones, B. D. and Sutterwala, F. S., Francisella tularensis live vaccine
strain folate metabolism and pseudouridine synthase gene mutants
modulate m acrophage caspase-1 activation. Infect. Immun. 2013. 81:
201–208.
77 Sampson, T. R., Napier, B. A., Schroeder, M. R., Louwen, R., Zhao, J.,
Chin, C. Y., Ratner, H. K. et al., A CRISPR-Cas system enhances envelope
integrity mediating antibiotic resistance and inflammasome evasion.
Proc. Natl. Acad. Sci. USA 2014. 111: 11163–11168.
78 Ge, J., Gong,Y.N., Xu, Y. and Shao, F., Preventing bacterial DNA release
and absent in melanoma 2 inflammasome activation by a Legionella
effector functioning in membrane trafficking. Proc. Natl. Acad. Sci. USA
2012. 109: 6193–6198.
79 Dotson,R.J., Rabadi, S. M., Westcott, E. L., Bradley, S., Catlett, S. V.,
Banik, S., Harton, J. A. et al., Repression of inflammasome by Francisella
tularensis during early stages of infection. J. Biol. Chem. 2013. 288: 23844–
23857.
80 Kanneganti, T. D., Central roles of NLRs and inflammasomes in viral
infection. Nat. Rev. Immunol. 2010. 10: 688–698.
81 Lupfer, C., Malik, A. and Kanneganti, T. D., Inflammasome control of
viral infection. Curr. Opin. Virol. 2015. 12: 38–46.
82 Reinholz, M., Kawakami, Y., Salzer, S.,
Kreuter, A., Dombrowski, Y.,
Koglin, S., Kresse, S. et al., HPV16 activates the AIM2 inflammasome in
keratinocytes. Arch. Dermatol. Res. 2013. 305: 723–732.
83 Zhen, J., Zhang, L., Pan, J., Ma, S., Yu, X., Li, X., Chen, S. et al.,AIM2
mediates inflammation-associated renal damage in hepatitis B virus-
associated glomerulonephritis by regulating caspase-1, IL-1beta, and
IL-18. Mediators Inflamm. 2014. 2014: 190860.
84 Wu,D.L., Xu, G. H., Lu,S.M., Ma,B.L., Miao, N. Z., Liu, X. B., Cheng, Y. P.
et al., Correlation of AIM2 expression in peripheral blood mononuclear
cells from humans with acute and chronic hepatitis B. Hum. Immunol.
2013. 74: 514–521.
85 Han, Y., Chen, Z., Hou, R., Yan, D., Liu, C., Chen, S., Li, X. et al.,Expres-
sion of AIM2 is correlated with increased inflammation in chronic hep-
atitis B patients. Virol. J. 2015. 12: 129.
86 Horan, K. A., Hansen, K., Jakobsen, M. R., Holm,C.K., Soby, S., Unter-
holzner, L., Thompson, M. et al., Proteasomal degradation of herpes
simplex virus capsids in macrophages releases DNA to the cytosol for
recognition by DNA sensors. J. Immunol. 2013. 190: 2311–2319.
87 Kanneganti, T. D., Ozoren, N., Body-Malapel, M.
, Amer, A., Park, J. H.,
Franchi, L., Whitfield, J. et al., Bacterial RNA and small antiviral com-
pounds activate caspase-1 through cryopyrin/Nalp3. Nature 2006. 440:
233–236.
88 Sander, L. E., Davis, M. J., Boekschoten, M. V., Amsen, D., Dascher, C.
C., Ryffel, B., Swanson, J. A. et al., Detection of prokaryotic mRNA signi-
fies microbial viability and promotes immunity. Nature 2011. 474: 385–
389.
89 Kailasan Vanaja, S., Rathinam, V. A., Atianand, M. K., Kalantari, P.,
Skehan, B., Fitzgerald, K. A. and Leong, J. M., Bacterial RNA:DNA hybrids
are activators of the NLRP3 inflammasome. Proc. Natl. Acad. Sci. USA
2014. 111: 7765–7770.
90 Ablasser, A., Bauernfeind, F., Hartmann, G., Latz, E., Fitzgerald, K. A.
and Hornung, V., RIG-I-dependent sensing of poly(dA:dT) through the
induction of an RNA polymerase III-transcribed RNA intermediate. Nat.
Immunol. 2009. 10: 1065–1072.
91 Chiu, Y. H., Macmillan, J. B. and Chen, Z. J., RNA polymerase III detects
cytosolic DNA and induces type I interferons through the RIG-I pathway.
Cell 2009. 138: 576–591.
92 Ekchariyawat, P., Hamel, R., Bernard, E., Wichit, S., Surasombatpat-
tana, P., Talignani, L., Thomas, F. et al., Inflammasome signaling path-
ways exert antiviral effect against Chikungunya virus in human dermal
fibroblasts. Infect. Genet. Evol. 2015.
32: 401–408.
93 Karki, R., Man, S. M., Malireddi, R. K., Gurung, P., Vogel, P., Lamkanfi,
M. and Kanneganti, T. D., Concerted activation of the AIM2 and NLRP3
inflammasomes orchestrates host protection against Aspergillus infec-
tion. Cell Host Microbe 2015. 17: 357–368.
94 Kalantari, P., DeOliveira, R. B., Chan, J., Corbett, Y., Rathinam, V., Stutz,
A., Latz, E. et al., Dual engagement of the NLRP3 and AIM2 inflamma-
somes by plasmodium-derived hemozoin and DNA during malaria. Cell
Rep. 2014. 6: 196–210.
95 Man, S. M., Zhu, Q., Zhu, L., Liu, Z., Karki, R., Malik, A., Sharma, D.
et al., Critical role for the DNA sensor AIM2 in stem cell proliferation
and cancer. Cell 2015. 162: 45–58.
96 Wilson, J. E., Petrucelli, A. S., Chen, L., Koblansky, A. A., Truax, A. D.,
Oyama, Y., Rogers, A. B. et al., Inflammasome-independent role of AIM2
in suppressing colon tumorigenesis via DNA-PK and Akt. Nat. Med. 2015.
21: 906–913.
97 DeYoung,K.L., Ray, M. E., Su, Y. A., Anzick,S.L., Johnstone, R. W., Tra-
pani, J. A.,
Meltzer, P. S. et al., Cloning a novel member of the human
interferon-inducible gene family associated with control of tumori-
genicity in a model of human melanoma. Oncogene 1997. 15: 453–457.
98 Dihlmann, S., Tao, S., Echterdiek, F., Herpel, E., Jansen, L., Chang-
Claude, J., Brenner, H. et al., Lack of absent in Melanoma 2 (AIM2)
expression in tumor cells is closely associated with poor survival in
colorectal cancer patients. Int. J. Cancer 2014. 135: 2387–2396.
99 Schulmann, K., Brasch,F.E., Kunstmann, E., Engel, C., Pagenstecher, C.,
Vogelsang, H., Kruger, S. et al., HNPCC-associated small bowel cancer:
clinical and molecular characteristics. Gastroenterology 2005. 128: 590–
599.
100 Woerner, S. M., Kloor, M., Schwitalle, Y., Youmans, H., Doeberitz, M.,
Gebert, J. and Dihlmann, S., The putative tumor suppressor AIM2 is fre-
quently affected by different genetic alterations in microsatellite unsta-
ble colon cancers. Genes Chromosomes Cancer 2007. 46: 1080–1089.
101 Kim, T. M., Laird, P. W. and Park, P. J., The landscape of microsatellite
instability in colorectal and endometrial cancer genomes. Cell 2013. 155:
858–868.
102 Ponomareva, L., Liu, H., Duan, X., Dickerson, E., Shen, H., Pan-
chanathan, R. and Choubey, D., AIM2, an IFN-inducible cytosolic DNA
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
280 Si Ming Man et al. Eur. J. Immunol. 2016. 46: 269–280
sensor, in the development of benign prostate hyperplasia and prostate
cancer. Mol. Cancer Res. 2013. 11: 1193–1202.
103 Wang,L.J., Hsu, C. W., Chen,C.C., Liang, Y., Chen,L.C., Ojcius, D. M.,
Tsang, N. M. et al., Interactome-wide analysis identifies end-binding
protein 1 as a crucial component for the speck-like particle formation
of activated absence in melanoma 2 (AIM2) inflammasomes. Mol. Cell.
Proteomics 2012. 11: 1230–1244.
104 Chen, L. C., Wang,L.J., Tsang, N. M., Ojcius, D. M., Chen, C. C., Ouyang,
C. N., Hsueh, C. et al., Tumour inflammasome-derived IL-1beta recruits
neutrophils and improves local recurrence-free survival in EBV-induced
nasopharyngeal carcinoma. EMBO Mol. Med. 2012. 4: 1276–1293.
105 Kondo, Y., Nagai, K., Nakahata, S., Saito, Y., Ichikawa, T., Suekane, A.,
Taki, T. et al., Overexpression of the DNA sensor proteins, absent in
melanoma 2 and interferon-inducible 16, contributes to tumorigenesis
of oral squamous cell carcinoma with p53 inactivation. Cancer Sci. 2012.
103: 782–790.
106 Kong, H., Wang, Y., Zeng, X., Wang, Z., Wang, H. and Xie, W., Differen-
tial expression of inflammasomes in lung cancer cell lines and tissues.
Tumour Biol. 2015. 36: 7501–7513.
107 Patsos, G., Germann, A., Gebert, J. and Dihlmann, S., Restoration of
absent in melanoma 2 (AIM2) induces G2/M cell cycle arrest and pro-
motes invasion of colorectal cancer cells. Int. J. Cancer 2010. 126: 1838–
1849.
108 Feng, J., Park, J.,
Cron, P., Hess, D. and Hemmings, B. A., Identification of
a PKB/Akt hydrophobic motif Ser-473 kinase as DNA-dependent protein
kinase. J. Biol. Chem. 2004. 279: 41189–41196.
109 Lu, D., Huang, J. and Basu, A., Protein kinase Cepsilon activates protein
kinase B/Akt via DNA-PK to protect against tumor necrosis factor-alpha-
induced cell death. J. Biol. Chem. 2006. 281: 22799–22807.
110 de Koning, H. D., Bergboer,J.G., van den Bogaard, E. H., van Vlijmen-
Willems, I. M., Rodijk-Olthuis, D., Simon, A., Zeeuwen, P. L. et al.,
Strong induction of AIM2 expression in human epidermis in acute
and chronic inflammatory skin conditions. Exp. Dermatol. 2012. 21:
961–964.
111 Yang,C.A., Huang, S. T. and Chiang, B. L., Sex-dependent differential
activation of NLRP3 and AIM2 inflammasomes in SLE macrophages.
Rheumatology 2015. 54: 324–331.
112 Kimkong, I., Avihingsanon, Y. and Hirankarn, N., Expression profile
of HIN200 in leukocytes and renal biopsy of SLE patients by real-time
RT-PCR. Lupus 2009. 18: 1066–1072.
113 Javierre, B. M., Fernandez, A. F., Richter, J., Al-Shahrour, F., Martin-
Subero, J. I., Rodriguez-Ubreva, J., Berdasco, M. et al., Changes in the
pattern of DNA methylation associate with twin discordance in sys-
temic lupus erythematosus. Genome Res. 2010. 20: 170–179.
114 Vanhove, W., Peeters, P. M., Staelens, D., Schraenen, A., Van der Goten,
J., Cleynen, I., De Schepper, S. et al., Strong upregulation of AIM2 and
IFI16 inflammasomes in the mucosa of patients with active inflamma-
tory bowel disease. Inflamm. Bowel Dis. 2015.
21: 2673–2682.
115 Lozano-Ruiz, B., Bachiller, V., Garcia-Martinez, I., Zapater, P., Gomez-
Hurtado, I., Moratalla, A., Gimenez, P. et al., Absent in melanoma 2 trig-
gers a heightened inflammasome response in ascitic fluid macrophages
of patients with cirrhosis. J. Hepatol. 2015. 62: 64–71.
116 Csak, T., Pillai, A., Ganz, M., Lippai, D., Petrasek, J., Park, J. K., Kodys, K.
et al., Both bone marrow-derived and non-bone marrow-derived cells
contribute to AIM2 and NLRP3 inflammasome activation in a MyD88-
dependent manner in dietary steatohepatitis. Liver Int. 2014. 34: 1402–
1413.
117 Adamczak, S. E., de Rivero Vaccari, J. P., Dale, G., Brand, F. J., 3rd.,
Nonner, D., Bullock, M. R., Dahl, G. P. et al., Pyroptotic neuronal cell
death mediated by the AIM2 inflammasome. J. Cereb. Blood Flow Metab.
2014. 34: 621–629.
118 Denes, A., Coutts, G., Lenart, N., Cruickshank, S. M., Pelegrin, P., Skin-
ner, J., Rothwell, N. et al., AIM2 and NLRC4 inflammasomes contribute
with ASC to acute brain injury independently of NLRP3. Proc. Natl. Acad.
Sci. USA 2015. 112: 4050–4055.
119 Kawane, K., Fukuyama, H., Kondoh, G., Takeda, J., Ohsawa, Y.,
Uchiyama, Y. and Nagata, S., Requirement of DNase II for definitive
erythropoiesis in the mouse fetal liver.
Science 2001. 292: 1546–1549.
120 Yoshida, H., Okabe, Y., Kawane, K., Fukuyama, H. and Nagata, S., Lethal
anemia caused by interferon-beta produced in mouse embryos carrying
undigested DNA. Nat. Immunol. 2005. 6: 49–56.
121 Kawane, K., Ohtani, M., Miwa, K., Kizawa, T., Kanbara, Y., Yoshioka,
Y., Yoshikawa, H. et al., Chronic polyarthritis caused by mammalian
DNA that escapes from degradation in macrophages. Nature 2006. 443:
998–1002.
122 Baum, R., Sharma, S., Carpenter, S., Li, Q. Z., Busto, P., Fitzgerald, K. A.,
Marshak-Rothstein, A. et al., Cutting edge: AIM2 and endosomal TLRs
differentially regulate arthritis and autoantibody production in DNase
II-deficient mice. J. Immunol. 2015. 194: 873–877.
123 Jakobs, C., Perner, S. and Hornung, V., AIM2 drives joint inflammation
in a self-DNA triggered model of chronic polyarthritis. PLoS One 2015.
10: e0131702.
124 Kaminski, J. J., Schattgen, S. A., Tzeng, T. C., Bode, C., Klinman, D.
M. and Fitzgerald, K. A., Synthetic oligodeoxynucleotides containing
suppressive TTAGGG motifs inhibit AIM2 inflammasome activation.
J. Immunol. 2013. 191: 3876–3883.
Abbreviations: ASC: apoptosis-associated speck-like protein containing
a carboxy-terminal caspase activation and recruitment domain · cGAS:
cyclic-GMP-AMP synthase · CSF: cerebrospinal fluid · GBP: guanylate-
binding protein · IRF: interferon-regulatory factor · LVS: live vaccine
strain · MCMV: mouse cytomegalovirus · SLE: systemic lupus erythe-
matosus
Full correspondence: Dr. Thirumala-Devi Kanneganti, Department of
Immunology, St. Jude Children’s Research Hospital MS #351, 262
Danny Thomas Place Memphis TN 38105-3678, USA
Fax: +1-(901) 595-5766
e-mail: Thirumala-Devi.Kanneganti@STJUDE.ORG
Received: 28/9/2015
Revised: 13/11/2015
Accepted: 26/11/2015
Accepted article online: 2/12/2015
C
2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.eji-journal.eu
... This inflammasome formation process activates caspase-1, which in turn cleaves pro-IL-1β and pro-IL-18, and activates the pore-forming protein gasdermin D (GSDMD). Next, GSDMD activation induces pyrotosis, an inflammatory form of cell death, and causes massive release of mature forms of IL-1β and IL-18 outside the cells (Man et al., 2016). ...
Article
Full-text available
Dendritic cells (DCs) are involved in the initiation and maintenance of immune responses against malignant cells by recognizing conserved pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) through pattern recognition receptors (PRRs). According to recent studies, tumor cell-derived DNA molecules act as DAMPs and are recognized by DNA sensors in DCs. Once identified by sensors in DCs, these DNA molecules trigger multiple signaling cascades to promote various cytokines secretion, including type I IFN, and then to induce DCs mediated antitumor immunity. As one of the potential attractive strategies for cancer therapy, various agonists targeting DNA sensors are extensively explored including the combination with other cancer immunotherapies or the direct usage as major components of cancer vaccines. Moreover, this review highlights different mechanisms through which tumor-derived DNA initiates DCs activation and the mechanisms through which the tumor microenvironment regulates DNA sensing of DCs to promote tumor immune escape. The contributions of chemotherapy, radiotherapy, and checkpoint inhibitors in tumor therapy to the DNA sensing of DCs are also discussed. Finally, recent clinical progress in tumor therapy utilizing agonist-targeted DNA sensors is summarized. Indeed, understanding more about DNA sensing in DCs will help to understand more about tumor immunotherapy and improve the efficacy of DC-targeted treatment in cancer.
... Only IFI16 and AIM2 can activate the inflammasome [12,13]. AIM2 and IFI16 have also shown the other same features, such as cytosolic dsDNA sensors, ASC (apoptosis-associated speck-like protein containing a CARD)dependent inflammasomes, regulating atherosclerosis, autoimmunity, tumorigenesis, and normal neuronal development [1,[14][15][16][17][18][19]. However, research shows that IFI16-β, a human mRNA variation of IFI16, inhibits the AIM2 inflammasome [20]. ...
Article
Full-text available
Interferon Gamma Inducible Protein 16 (IFI16) belongs to the HIN-200 protein family and is pivotal in immunological responses. Serving as a DNA sensor, IFI16 identifies viral and aberrant DNA, triggering immune and inflammatory responses. It is implicated in diverse cellular death mechanisms, such as pyroptosis, apoptosis, and necroptosis. Notably, these processes are integral to the emergent concept of PANoptosis, which encompasses cellular demise and inflammatory pathways. Current research implies a significant regulatory role for IFI16 in PANoptosis, particularly regarding cardiac pathologies. This review delves into the complex interplay between IFI16 and PANoptosis in heart diseases, including atherosclerosis, myocardial infarction, heart failure, and diabetic cardiomyopathy. It synthesizes evidence of IFI16’s impact on PANoptosis, with the intention of providing novel insights for therapeutic strategies targeting heart diseases.
... During HSV-1 infection, the NLRP3 (NLR family pyrin domain containing 3), AIM 2 (Absent-in-melanoma-2), IFI16 (Interferon-inducible protein 16), and RIG-1 (Retinoic acid-inducible gene I)-dependent inflammasomes are known to be activated [20][21][22][23][24][25][26][27][28]. Specifically, AIM 2 is a sensor that can detect double-stranded DNA at the cytoplasmic level through its hematopoietic interferon-inducible nuclear HIN (hematopoietic interferon-inducible nuclear domain): its activation leads to the formation of the AIM 2 inflammasome through the oligomerization of AIM 2 protein sensor with ASC and pro-caspase-1, resulting in the activation of caspase-1 (p30 auto-clivated to p20 and p10) and, consequently, the release of IL-1β and IL-18 ( Figure 1) [29][30][31][32][33]. ...
Article
Full-text available
Although Herpes simplex virus type 1 (HSV-1) has been deeply studied, significant gaps remain in the fundamental understanding of HSV-host interactions: our work focused on studying the Infected Cell Protein 27 (ICP27) as an inhibitor of the Absent-in-melanoma-2 (AIM 2) inflammasome pathway, leading to reduced pro-inflammatory cytokines that influence the activation of a protective innate immune response to infection. To assess the inhibition of the inflammasome by the ICP27, hTert-immortalized Retinal Pigment Epithelial cells (hTert-RPE 1) infected with HSV-1 wild type were compared to HSV-1 lacking functional ICP27 (HSV-1∆ICP27) infected cells. The activation of the inflammasome by HSV-1∆ICP27 was demonstrated by quantifying the gene and protein expression of the inflammasome constituents using real-time PCR and Western blot. The detection of the cleavage of the pro-caspase-1 into the active form was performed by using a bioluminescent assay, while the quantification of interleukins 1β (IL-1β) and 18 (IL-18)released in the supernatant was quantified using an ELISA assay. The data showed that the presence of the ICP27 expressed by HSV-1 induces, in contrast to HSV-1∆ICP27 vector, a significant downregulation of AIM 2 inflammasome constituent proteins and, consequently, the release of pro-inflammatory interleukins into the extracellular environment reducing an effective response in counteracting infection.
... However, dsDNA less than 80 bp in length is a poor activator of the AIM2 inflammasome. AIM2 can recognize self-DNA and activate inflammasome complexes, but its cytosolic location limits the recognition of self-DNA under steady-state conditions [101]. Nonetheless, in situations where self-DNA is inefficiently cleared from the extracellular milieu or improperly degraded in the phagolysosomal compartment, it can access cytosolic compartments, overexpress genes encoding type I IFNs and associated genes, and activate AIM2 to drive inflammation [102]. ...
Article
Full-text available
Inflammasomes are complex platforms for the cleavage and release of inactivated IL-1β and IL-18 cytokines that trigger inflammatory responses against damage-associated molecular patterns (DAMPs) or pathogen-associated molecular patterns (PAMPs). Gut microbiota plays a pivotal role in maintaining gut homeostasis. Inflammasome activation needs to be tightly regulated to limit aberrant activation and bystander damage to the host cells. Several types of inflammasomes, including Node-like receptor protein family (e.g., NLRP1, NLRP3, NLRP6, NLRP12, NLRC4), PYHIN family, and pyrin inflammasomes, interact with gut microbiota to maintain gut homeostasis. This review discusses the current understanding of how inflammasomes and microbiota interact, and how this interaction impacts human health. Additionally, we introduce novel biologics and antagonists, such as inhibitors of IL-1β and inflammasomes, as therapeutic strategies for treating gastrointestinal disorders when inflammasomes are dysregulated or the composition of gut microbiota changes.
... As the best-characterized member of this protein family is AIM2 (Absent In Melanoma 2), they are also known as AIM2-like receptors (ALRs). Human AIM2 acts as a cytosolic immune sensor that triggers inflammasome activation upon detection of pathogen-derived double-stranded DNAs [4][5][6]. Humans encode three additional PYHIN proteins: γ-IFN-inducible protein 16 (IFI16), IFN-inducible protein X (IFIX), also known as Pyrin and HIN domain-containing protein 1 (PYHIN1), and Myeloid Nuclear Differentiation Antigen (MNDA) [7]. All four human PYHIN proteins contain an N-terminal Pyrin domain (PYD) and one or two C-terminal hematopoietic interferon-inducible nuclear proteins with a 200-amino-acid repeat domain (HIN200). ...
Article
Full-text available
PYHIN proteins are only found in mammals and play key roles in the defense against bacterial and viral pathogens. The corresponding gene locus shows variable deletion and expansion ranging from 0 genes in bats, over 1 in cows, and 4 in humans to a maximum of 13 in mice. While initially thought to act as cytosolic immune sensors that recognize foreign DNA, increasing evidence suggests that PYHIN proteins also inhibit viral pathogens by more direct mechanisms. Here, we examined the ability of all 13 murine PYHIN proteins to inhibit HIV-1 and murine leukemia virus (MLV). We show that overexpression of p203, p204, p205, p208, p209, p210, p211, and p212 strongly inhibits production of infectious HIV-1; p202, p207, and p213 had no significant effects, while p206 and p214 showed intermediate phenotypes. The inhibitory effects on infectious HIV-1 production correlated significantly with the suppression of reporter gene expression by a proviral Moloney MLV-eGFP construct and HIV-1 and Friend MLV LTR luciferase reporter constructs. Altogether, our data show that the antiretroviral activity of PYHIN proteins is conserved between men and mice and further support the key role of nuclear PYHIN proteins in innate antiviral immunity.
Article
Full-text available
Absent in melanoma 2 (AIM2) is implicated in neuroinflammation. Here, we explored the prognostic significance of serum AIM2 in human aneurysmal subarachnoid hemorrhage (aSAH). We conducted a consecutive enrollment of 127 patients, 56 of whom agreed with blood-drawings not only at admission but also at days 1, 2, 3, 5, 7 and 10 days after aSAH. Serum AIM2 levels of patients and 56 healthy controls were measured. Disease severity was assessed using the modified Fisher scale (mFisher) and World Federation of Neurological Surgeons Scale (WFNS). Neurological outcome at poststroke 90 days was evaluated via the modified Rankin Scale (mRS). Univariate analysis and multivariate analysis were sequentially done to ascertain relationship between serum AIM2 levels, severity, delayed cerebral ischemia (DCI) and 90-day poor prognosis (mRS scores of 3–6). Patients, in comparison to controls, had a significant elevation of serum AIM2 levels at admission and at days 1, 2, 3, 5, 7 and 10 days after aSAH, with the highest levels at days 1, 2, 3 and 5. AIM2 levels were independently correlated with WFNS scores and mFisher scores. Significantly higher serum AIM2 levels were detected in patients with a poor prognosis than in those with a good prognosis, as well as in patients with DCI than in those without DCI. Moreover, serum AIM2 levels independently predicted a poor prognosis and DCI, and were linearly correlated with their risks. Using subgroup analysis, there were no significant interactions between serum AIM2 levels and age, gender, hypertension and so on. There were substantially high predictive abilities of serum AIM2 for poor prognosis and DCI under the receiver operating characteristic curve. The combination models of DCI and poor prognosis, in which serum AIM2, WFNS scores and mFisher scores were incorporated, showed higher discriminatory efficiencies than anyone of the preceding three variables. Moreover, the models are delineated using the nomogram, and performed well under the calibration curve and decision curve. Serum AIM2 levels, with a substantial enhancement during early phase after aSAH, are closely related to bleeding severity, poor 90-day prognosis and DCI of patients, substantializing serum AIM2 as a potential prognostic biomarker of aSAH.
Article
Duchenne Muscular Dystrophy (DMD) is a progressive and fatal neuromuscular disease. Cycles of myofibre degeneration and regeneration are hallmarks of the disease where immune cells infiltrate to repair damaged skeletal muscle. Benfotiamine is a lipid soluble precursor to thiamine, shown clinically to reduce inflammation in diabetic related complications. We assessed whether benfotiamine administration could reduce inflammation related dystrophic pathology. Benfotiamine (10 mg/kg/day) was fed to male mdx mice (n = 7) for 15 weeks from 4 weeks of age. Treated mice had an increased growth weight (5–7 weeks) and myofibre size at treatment completion. Markers of dystrophic pathology (area of damaged necrotic tissue, central nuclei) were reduced in benfotiamine mdx quadriceps. Grip strength was increased and improved exercise capacity was found in mdx treated with benfotiamine for 12 weeks, before being placed into individual cages and allowed access to an exercise wheel for 3 weeks. Global gene expression profiling (RNAseq) in the gastrocnemius revealed benfotiamine regulated signalling pathways relevant to dystrophic pathology (Inflammatory Response, Myogenesis) and fibrotic gene markers (Col1a1, Col1a2, Col4a5, Col5a2, Col6a2, Col6a2, Col6a3, Lum) towards wildtype levels. In addition, we observed a reduction in gene expression of inflammatory gene markers in the quadriceps (Emr1, Cd163, Cd4, Cd8, Ifng). Overall, these data suggest that benfotiamine reduces dystrophic pathology by acting on inflammatory and fibrotic gene markers and signalling pathways. Given benfotiamine’s excellent safety profile and current clinical use, it could be used in combination with glucocorticoids to treat DMD patients.
Article
Full-text available
Interferon (IFN)-inducible Guanylate Binding Proteins (GBPs) mediate cell-autonomous host resistance to bacterial pathogens and promote inflammasome activation. The prevailing model postulates that these two GBP-controlled activities are directly linked through GBP-dependent vacuolar lysis. It was proposed that rupture of pathogen-containing vacuoles (PVs) by GBPs destroyed the microbial refuge and simultaneously contaminated the host cell cytosol with microbial activators of inflammasomes. Here, we demonstrate that GBP-mediated host resistance and GBP-mediated inflammatory responses can be uncoupled. We show that PVs formed by the rodent pathogen Chlamydia muridarum, so-called inclusions, remain free of GBPs and that C. muridarum is impervious to GBP-mediated restrictions on bacterial growth. Although GBPs neither bind to C. muridarum inclusions nor restrict C. muridarum growth, we find that GBPs promote inflammasome activation in C. muridarum-infected macrophages. We demonstrate that C. muridarum infections induce GBP-dependent pyroptosis through both caspase-11-dependent noncanonical and caspase-1-dependent canonical inflammasomes. Amongst canonical inflammasomes we find that C. muridarum and the human pathogen Chlamydia trachomatis not only activate NLRP3, as previously reported, but also AIM2. Our data show that GBPs support fast-kinetics processing and secretion of IL-1β and IL-18 by the NLRP3 inflammasome but are dispensable for the secretion of the same cytokines at later times post-infection. Because IFNγ fails to induce IL-1β transcription, GBP-dependent fast-kinetics inflammasome activation can drive the preferential processing of constitutively expressed IL-18 in IFNγ-primed macrophages in the absence of prior TLR stimulation. Together, our results reveal that GBPs control the kinetics of inflammasome activation and thereby shape macrophage responses to Chlamydia infections.
Article
Full-text available
Inflammatory caspases (caspase-1, -4, -5 and -11) are critical for innate defences. Caspase-1 is activated by ligands of various canonical inflammasomes, and caspase-4, -5 and -11 directly recognize bacterial lipopolysaccharide, both of which trigger pyroptosis. Despite the crucial role in immunity and endotoxic shock, the mechanism for pyroptosis induction by inflammatory caspases is unknown. Here we identify gasdermin D (Gsdmd) by genome-wide clustered regularly interspaced palindromic repeat (CRISPR)-Cas9 nuclease screens of caspase-11- and caspase-1-mediated pyroptosis in mouse bone marrow macrophages. GSDMD-deficient cells resisted the induction of pyroptosis by cytosolic lipopolysaccharide and known canonical inflammasome ligands. Interleukin-1β release was also diminished in Gsdmd(-/-) cells, despite intact processing by caspase-1. Caspase-1 and caspase-4/5/11 specifically cleaved the linker between the amino-terminal gasdermin-N and carboxy-terminal gasdermin-C domains in GSDMD, which was required and sufficient for pyroptosis. The cleavage released the intramolecular inhibition on the gasdermin-N domain that showed intrinsic pyroptosis-inducing activity. Other gasdermin family members were not cleaved by inflammatory caspases but shared the autoinhibition; gain-of-function mutations in Gsdma3 that cause alopecia and skin defects disrupted the autoinhibition, allowing its gasdermin-N domain to trigger pyroptosis. These findings offer insight into inflammasome-mediated immunity/diseases and also change our understanding of pyroptosis and programmed necrosis.
Article
Full-text available
Intracellular lipopolysaccharide from Gram-negative bacteria including Escherichia coli, Salmonella typhimurium, Shigella flexneri and Burkholderia thailandensis activates mouse caspase-11, causing pyroptotic cell death, interleukin-1β processing, and lethal septic shock. How caspase-11 executes these downstream signalling events is largely unknown. Here we show that gasdermin D is essential for caspase-11-dependent pyroptosis and interleukin-1β maturation. A forward genetic screen with ethyl-N-nitrosourea-mutagenized mice links Gsdmd to the intracellular lipopolysaccharide response. Macrophages from Gsdmd(-/-) mice generated by gene targeting also exhibit defective pyroptosis and interleukin-1β secretion induced by cytoplasmic lipopolysaccharide or Gram-negative bacteria. In addition, Gsdmd(-/-) mice are protected from a lethal dose of lipopolysaccharide. Mechanistically, caspase-11 cleaves gasdermin D, and the resulting amino-terminal fragment promotes both pyroptosis and NLRP3-dependent activation of caspase-1 in a cell-intrinsic manner. Our data identify gasdermin D as a critical target of caspase-11 and a key mediator of the host response against Gram-negative bacteria.
Article
Full-text available
AIM2 recognizes foreign dsDNA and assembles into the inflammasome, a filamentous supramolecular signalling platform required to launch innate immune responses. We show here that the pyrin domain of AIM2 (AIM2(PYD)) drives both filament formation and dsDNA binding. In addition, the dsDNA-binding domain of AIM2 also oligomerizes and assists in filament formation. The ability to oligomerize is critical for binding dsDNA, and in turn permits the size of dsDNA to regulate the assembly of the AIM2 polymers. The AIM2(PYD) oligomers define the filamentous structure, and the helical symmetry of the AIM2(PYD) filament is consistent with the filament assembled by the PYD of the downstream adaptor ASC. Our results suggest that the role of AIM2(PYD) is not autoinhibitory, but generating a structural template by coupling ligand binding and oligomerization is a key signal transduction mechanism in the AIM2 inflammasome.
Article
Full-text available
Mice lacking DNase II display a polyarthritis-like disease phenotype that is driven by translocation of self-DNA into the cytoplasm of phagocytic cells, where it is sensed by pattern recognition receptors. While pro-inflammatory gene expression is non-redundantly linked to the presence of STING in these mice, the contribution of the inflammasome pathway has not been explored. To this end, we studied the role of the DNA-sensing inflammasome receptor AIM2 in this self-DNA driven disease model. Arthritis-prone mice lacking AIM2 displayed strongly decreased signs of joint inflammation and associated histopathological findings. This was paralleled with a reduction of caspase-1 activation and pro-inflammatory cytokine production in diseased joints. Interestingly, systemic signs of inflammation that are associated with the lack of DNase II were not dependent on AIM2. Taken together, these data suggest a tissue-specific role for the AIM2 inflammasome as a sensor for endogenous DNA species in the course of a ligand-dependent autoinflammatory condition.
Article
Full-text available
The absent in melanoma 2 (AIM2), a cytosolic dsDNA inflammasome, can be activated by viral DNA to trigger caspase-1. Its role in immunopathology of chronic hepatitis B and C virus (HBV, HCV) infection is still largely unclear. In this study, the expression AIM2, and its downstream cytokines, caspase-1, IL-18 and IL-1β, in liver tissue of patients with chronic hepatitis B and C (CHB, CHC) were investigated. A total of 70 patients diagnosed with chronic hepatitis were enrolled, including 47 patients with CHB and 23 patients with CHC. A liver biopsy was taken from each patient, and immunohistochemistry was used to detect the expression of AIM2 and inflammatory factors caspase-1, IL-18, and IL-1β in the biopsy specimens. The relationship between AIM2 expression and these inflammatory factors was analyzed. The expression of AIM2 in CHB patients (89.4 %) was significantly higher than in CHC patients (8.7 %), and among the CHB patients, the expression of AIM2 was significantly higher in the high HBV replication group (HBV DNA ≥ 1 × 10(5)copies/mL) than in the low HBV replication group (HBV DNA < 1 × 10(5)copies/mL). The expression of AIM2 was also correlated with HBV-associated inflammatory activity in CHB patients statistically. Additionally, AIM2 levels were positively correlated with the expression of caspase-1, IL-1β and IL-18 in CHB patients, which implied that the AIM2 expression is directly correlated with the inflammatory activity associated with CHB. AIM2 upregulation may be a component of HBV immunopathology. The underlying mechanism and possible signal pathway warrant further study.
Article
Inflammasomes are protein complexes that promote caspase activation, resulting in processing of IL-1β and cell death, in response to infection and cellular stresses. Inflammasomes have been anticipated to contribute to autoimmunity. The New Zealand Black (NZB) mouse develops anti-erythrocyte Abs and is a model of autoimmune hemolytic anemia. These mice also develop anti-nuclear Abs typical of lupus. In this article, we show that NZB macrophages have deficient inflammasome responses to a DNA virus and fungal infection. Absent in melanoma 2 (AIM2) inflammasome responses are compromised in NZB by high expression of the AIM 2 antagonist protein p202, and consequently NZB cells had low IL-1β output in response to both transfected DNA and mouse CMV infection. Surprisingly, we also found that a second inflammasome system, mediated by the NLR family, pyrin domain containing 3 (NLRP3) initiating protein, was completely lacking in NZB cells. This was due to a point mutation in an intron of the Nlrp3 gene in NZB mice, which generates a novel splice acceptor site. This leads to incorporation of a pseudoexon with a premature stop codon. The lack of full-length NLRP3 protein results in NZB being effectively null for Nlrp3, with no production of bioactive IL-1β in response to NLRP3 stimuli, including infection with Candida albicans. Thus, this autoimmune strain harbors two inflammasome deficiencies, mediated through quite distinct mechanisms. We hypothesize that the inflammasome deficiencies in NZB alter the interaction of the host with both microflora and pathogens, promoting prolonged production of cytokines that contribute to development of autoantibodies.
Article
In response to infections and tissue damage, ASC-containing inflammasome protein complexes are assembled that promote caspase-1 activation, IL-1β and IL-18 processing and release, pyroptosis, and the release of ASC particles. However, excessive or persistent activation of the inflammasome causes inflammatory diseases. Therefore, a well-balanced inflammasome response is crucial for the maintenance of homeostasis. We show that the PYD-only protein POP1 inhibited ASC-dependent inflammasome assembly by preventing inflammasome nucleation, and consequently interfered with caspase-1 activation, IL-1β and IL-18 release, pyroptosis, and the release of ASC particles. There is no mouse ortholog for POP1, but transgenic expression of human POP1 in monocytes, macrophages, and dendritic cells protected mice from systemic inflammation triggered by molecular PAMPs, inflammasome component NLRP3 mutation, and ASC danger particles. POP1 expression was regulated by TLR and IL-1R signaling, and we propose that POP1 provides a regulatory feedback loop that shuts down excessive inflammatory responses and thereby prevents systemic inflammation. Copyright © 2015 Elsevier Inc. All rights reserved.