ArticlePDF Available

Caspase-8 mediates caspase-1 processing and innate immune defense in response to bacterial blockade of NF- B and MAPK signaling

Authors:

Abstract and Figures

Significance Pathogenic organisms express virulence factors that can inhibit immune signaling pathways. Thus, the immune system is faced with the challenge of eliciting an effective inflammatory response to pathogens that actively suppress inflammation. The mechanisms that regulate this response are largely undefined. The Yersinia virulence factor YopJ blocks NF-κB and MAPK signaling, resulting in reduced cytokine production and target cell death. Here, we find that caspase-8, RIPK1, and FADD are required for YopJ-induced cell death and show that mice lacking caspase-8 are severely susceptible to Yersinia infection and have defective proinflammatory cytokine production. These findings highlight a possible mechanism of immune defense that can overcome pathogen inhibition of cell-intrinsic proinflammatory immune responses.
Content may be subject to copyright.
Caspase-8 and RIP kinases regulate bacteria-induced
innate immune responses and cell death
Dan Weng
a
, Robyn Marty-Roix
a
, Sandhya Ganesan
a
, Megan K. Proulx
b
, Gregory I. Vladimer
a
, William J. Kaiser
c
,
Edward S. Mocarski
c
, Kimberly Pouliot
a
, Francis Ka-Ming Chan
d
, Michelle A. Kelliher
e
, Phillip A. Harris
f
, John Bertin
f
,
Peter J. Gough
f
, Dmitry M. Shayakhmetov
g
, Jon D. Goguen
b
, Katherine A. Fitzgerald
a,h
, Neal Silverman
a
,
and Egil Lien
a,h,1
a
Program in Innate Immunity, Division of Infectious Diseases and Immunology, Department of Medicine,
b
Department of Microbiology and Physiological
Systems,
d
Department of Cancer Biology, and
e
Department of Pathology, University of Massachusetts Medical School, Worcester, MA 01605;
c
Department of
Microbiology and Immunology, Emory Vaccine Center, Emory University School of Medicine, Atlanta, GA 30322;
f
Pattern Recognition Receptor Discovery
Performance Unit, Immuno-inflammation Therapeutic Area, GlaxoSmithKline, Collegeville, PA 19426;
g
Lowance Center for Human Immunology, Departments
of Pediatrics and Medicine, Emory University, Atlanta, GA 30322; and
h
Centre of Molecular Inflammation Research, Department of Cancer Research and
Molecular Medicine, Norwegian University of Science and Technology, 7491 Trondheim, Norway
Edited by Ruslan Medzhitov, Yale University School of Medicine, New Haven, CT, and approved April 1, 2014 (received for review February 25, 2014)
A number of pathogens cause host cell death upon infection, and
Yersinia pestis, infamous for its role in large pandemics such as the
Black Deathin medieval Europe, induces considerable cytotoxic-
ity. The rapid killing of macrophages induced by Y. pestis, depen-
dent upon type III secretion system effector Yersinia outer protein
J (YopJ), is minimally affected by the absence of caspase-1, cas-
pase-11, Fas ligand, and TNF. Caspase-8 is known to mediate apo-
ptotic death in response to infection with several viruses and to
regulate programmed necrosis (necroptosis), but its role in bacte-
rially induced cell death is poorly understood. Here we provide
genetic evidence for a receptor-interacting protein (RIP) kinase
caspase-8-dependent macrophage apoptotic death pathway after
infection with Y. pestis, influenced by Toll-like receptor 4-TIR-do-
main-containing adapter-inducing interferon-β(TLR4-TRIF). Inter-
estingly, macrophages lacking either RIP1, or caspase-8 and RIP3,
also had reduced infection-induced production of IL-1β, IL-18, TNF,
and IL-6; impaired activation of the transcription factor NF-κB; and
greatly compromised caspase-1 processing. Cleavage of the proform
of caspase-1 is associated with triggering inflammasome activity,
which leads to the maturation of IL-1βand IL-18, cytokines impor-
tant to host responses against Y. pestis and many other infectious
agents. Our results identify a RIP1caspase-8/RIP3-dependent cas-
pase-1 activation pathway after Y. pestis challenge. Mice defective
in caspase-8 and RIP3 were also highly susceptible to infection and
displayed reduced proinflammatory cytokines and myeloid cell death.
We propose that caspase-8 and the RIP kinases are key regulators of
macrophage cell death, NF-κB and inflammasome activation, and host
resistance after Y. pestis infection.
The causative agent of plague, Yersinia pestis is well known to
cause significant cell death upon infection (13). Like the
activation of inflammatory pathways to produce cytokines, trig-
gering cell death pathways is a common response of the mam-
malian immune system to infection. Death of immune cells can
eliminate the replication niche of pathogens found within those
cells, thus inhibiting the proliferation of the pathogens and ex-
posing them to bactericidal mechanisms (4). Conversely, elimi-
nation of key immune cells can diminish the ability of those cells
to respond to infection. Multiple host and microbial factors control
cell death pathways (5). Caspase-8dependent apoptosis, receptor
interacting protein-1 (RIP1)- and RIP3-dependent necroptosis, and
caspase-1/caspase-11dependent pyroptosis constitute major modes
of regulated cell death during infection (5, 6). Several viruses seem to
induce caspase-8dependent apoptosis (7). Caspase-8 has also been
suggested to have additional functions, such as inhibiting necroptosis
(79) and modulation of NF-κBactivationinTandBcells(10).
Signaling to the transcription factor NF-κB controls the transcription
of cytokines such as IL-6, TNF, pro-IL-1β, and pro-IL-18, and
stimulates cell survival. Y. pestis can induce cell death in macrophages
and dendritic cells via the type III secretion system (T3SS) effector
Yersinia outer protein J (YopJ; YopP in Yersinia enterocolitica), al-
though it is unclear whether this is entirely by apoptosis (11, 12). All
human-pathogenic Yersiniae (Y. pestis,Yersinia pseudotuberculosis,
and Y. enterocolitica) harbor cytotoxic properties toward host cells,
and YopJ production is associated with cell death in vivo and in
vitro (1316). YopJ-mediated inhibition of NF-κB by acetylation of
Inhibitor of κBKinaseβ(IKKβ), MAP kinase kinases, and TAK1
may modulate macrophage death via effects on inflammatory and
prosurvival signals (2, 1721). Inflammasome activation, culminat-
ing in the activation and processing of caspase-1, leads to the pro-
duction of IL-18 and IL-1β, key inflammatory cytokines and
antibacterial defenses, but can also be associated with caspase-1
dependent pyroptotic cell death (22). YopJ also participates in
inflammasome activation (16, 23), leading to a host immune re-
sponse. Thus, this single bacterial effector may induce both pro-
tective and harmful effects for the host. In the present study we
investigated the mechanisms for Y. pestis-induced cell death, NF-κB
activation, and triggering of inflammasome activation.
Results and Discussion
Yersinia Induces Cell Death via RIP1, Caspase-8, and RIP3. Viable
Y. pestis KIM5 can induce rapid cell death via YopJ (Fig. S1A).
Rapid death in bone marrow-derived macrophages (BMDMs) is
induced in a YopJ-dependent manner by Y. pestis or Y. pseudo-
tuberculosis temperature-shifted from 26 °C to 37 °C (Fig. S1 A
Significance
Receptor-interacting protein-1 (RIP1) kinase and caspase-8 are
important players in activation of apoptotic pathways. Here
we show that RIP1, caspase-8, and RIP3 contribute to infection-
induced macrophage cell death and also are required for acti-
vation of transcription factor NF-κB and caspase-1 upon in-
fection with the bacterial pathogen Yersinia pestis, the causative
agent of plague. Mice lacking caspase-8 and RIP3 are also very
susceptible to bacterial infection. This suggests that RIP1, cas-
pase-8, and RIP3 are key molecules with multiple roles in innate
immunity during bacterial challenge.
Author contributions: D.W., R.M.-R., and E.L. designed research; D.W., R.M.-R., G.I.V., and
E.L. performed research; S.G., M.K.P., W.J.K., E.S.M., K.P., F.K.-M.C., M.A.K., P.A.H., J.B.,
P.J.G., D.M.S., J.D.G., K.A.F., and N.S.contributed new reagents/analytic tools; D.W., R.M.-R.,
K.A.F., N.S., and E.L. analyzed data; and D.W., R.M.-R., J.D.G., K.A.F., N.S., and E.L. wrote
the paper.
Conflict of interest statement: E.L. and J.D.G. have a patent application on the use of
modified bacteria as used in vaccines. P.A.H., J.B., and P.J.G. are employees and share-
holders of GlaxoSmithKline.
This article is a PNAS Direct Submission.
1
To whom correspondence should be addressed. E-mail: Egil.Lien@umassmed.edu.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1403477111/-/DCSupplemental.
www.pnas.org/cgi/doi/10.1073/pnas.1403477111 PNAS
|
May 20, 2014
|
vol. 111
|
no. 20
|
73917396
IMMUNOLOGY
and B), a condition that mimics the temperature change associated
with infection via a fleabite. In addition to arming the T3SS, the
temperature shift ensures the initial presence of some TLR4-
stimulatory LPS (Fig. S1C) (24). Although caspase-1 is activated
by Y. pestis (1, 25), macrophage death was independent of cas-
pase-1/caspase-11, suggesting nonpyroptotic cytotoxicity (Fig. S2
Aand B). Death was unaffected by Fas ligand (FasL) or TNF,
indicating that those death receptor-mediated mechanisms are
not involved; and independent of inflammasome-related NOD-
like receptors (NLRs), the RNA-dependent protein kinase
(PKR), the inflammasome adaptor apoptosis-associated speck-
like protein containing a CARD (ASC), IL-1β, and IL-18 (Fig.
S2 CG) (1). Caspase-8 is a key enzyme in cell death induced by
some viruses (26). Caspase-8 deficiency results in embryonic le-
thality, but mice deficient in both caspase-8 and RIP3 [RIP3
/
caspase-8
/
mice, double knockout (dKO)] are rescued. These
data indicate a vital role for caspase-8 in suppressing necroptosis
by targeting a component of the RIP3 pathway (8, 9). Macro-
phages from RIP3
/
caspase-8
/
mice were remarkably resistant
to cell death induced by Y. pestis and Y. pseudotuberculosis, but not
by Salmonella, which induces pyroptotic death (4), or with the
NLRP3 inflammasome-specific trigger nigericin (Fig. 1 Aand B).
YopJ-induced death is likely not necroptosis because RIP3-
deficient cells are not protected (Fig. 1 Aand B). Electron
microscopy revealed that macrophages infected with Y. pestis
displayed features consistent with apoptotic death, such as
membrane blebbing and nuclear condensation and fragmen-
tation. These effects were absent in visibly infected dKO cells (Fig.
1C). Moreover, infection of macrophages with Y. pestis led to
DNA fragmentation patterns typically associated with apopto-
sis, and this was blocked by zVAD pan-caspase inhibition (Fig. S3A).
Taken together, our data strongly suggest that Yersinia induces
rapid macrophage death by apoptosis via caspase-8.
RIP3-mediated necroptotic death requires RIP1 (2729), a
serine/threonine kinase that canalso contribute to NF-κB signaling
(30) and apoptosis. RIP1
/
mice die shortly after birth (31), but
fetal liver macrophages from RIP1
/
mice, in contrast to RIP3
/
macrophages, displayed a rescue from death induced by Y. pestis
(Fig. 2A) and DNA laddering (Fig. S3C), suggesting that RIP1
activity contributes to apoptotic cell death upon infection, likely
mediated by the induction of caspase-8 enzymatic activity and
cleavage of procaspase-8 that precedes cell death (Figs. 1D,and2A,
B,E,andFand Fig. S3 Dand E). RIP1
/
macrophages were also
protected from necroptotic cell death induced by heat-killed
KIM5 plus zVAD or LPS plus zVAD (Fig. S3F). Potent and speci-
fic inhibitors of RIP3 (32) or RIP1 [GlaxoSmithKline (GSK): P.A.H.,
J.B., P.J.G.] kinase activity have recently been identified. The
RIP1 inhibitor GSK963, but not inactive enantiomer GSK962,
blocks Y. pestis-induced cell death (Fig. 2Cand Fig. S3B) and
caspase-8 activity (Fig. 2D). In addition to the genetic and phar-
macological interactions between RIP1 and caspase-8, we found
that RIP1 biochemically interacted with caspase-8 after Y. pestis
challenge (Fig. 2E). Cell death, cleavage of procaspase-8, and
enzymatic activity were partially reduced in the absence of TLR4
and TRIF, but not MyD88 (Fig. S4). Reduced death was also seen
for bacteria grown at 37 °C and Y. pestis-EcLpxL, which consti-
tutively generates a hexa-acylated LPS (Fig. S5 Aand B). TLR4
signaling seems to enhance early caspase-8mediated effects by Y.
pestis YopJ, similar to those proposed for Y. enterocolitica YopP
(3335). Cell death induced by Y. pestis grown at 37 °C was
inhibited by the presence of CaF1 capsule protein (Fig. S5 Cand
D), suggesting that the capsule prevented close contact between
bacteria and host cells needed for T3SS effects.
The targeted deletion of caspase-8 in myeloid cells [condi-
tional KO (cKO) caspase-8
fl/fl
LysM cre
+/+
generated by D.M.S.;
Fig. S6 Aand B] had little effect on Y. pestis-induced macrophage
death (Fig. 2C). Although the generation of other mice with
defects in caspase-8 in macrophages has been reported (36), our
caspase-8 cKO BMDM appeared healthy in culture and did not
display increased cell death in the presence or absence of infection
(Fig. 2C). Blockade of RIP3 kinase activity with GSK872 strongly
reduced macrophage death in the absence of caspase-8, suggesting
that deletion of caspase-8, or caspase inhibition by zVAD (Fig. 2C
and Fig. S6C), may promote necroptosis by RIP3, presumably
influenced by reduced Y. pestis-induced cleavage of RIP1 in the
absence of caspase-8 (Fig. 2F).
Cleavage and activation of the downstream apoptotic execu-
tioner caspase-3 was also dependent upon YopJ and caspase-8
RIP3 (Fig. S6D). The caspase-8RIP3 pathway also influenced
death induced by Y. enterocolitica but not by Salmonella or
Pseudomonas, which also harbor a T3SS (Fig. S6E). Thus, all
human-pathogenic Yersiniae, but not all bacteria containing a
T3SS, trigger cell death via the same pathway. Our results pro-
vide an explanation for how Yersinia induces macrophage cell
death via caspase-8 and RIP kinases. In this model, caspase-8
dependent apoptosis represents the default, whereas caspase-8
absence may lead to RIP3-dependent necroptosis. RIP1 has a key
upstream role for both modes of death, perhaps influenced by its
ability to direct apoptosis under conditions of cIAP1 depletion
(37) as seen with Y. pestis (Fig. S6F).
Effects on NF-κB Activity. Caspase-8 has also been suggested to
regulate NF-κB activity (10, 38, 39). We found a reduction in TNF
and IL-6 release, and pro-IL-1βexpression, all controlled by NF-
κB, in RIP1 KO, caspase-8 cKO, and RIP3
/
caspase-8
/
, but not
in RIP3
/
macrophages upon infection or LPS treatment (Fig. 3
AF). However, cytokine production by the TLR2 ligand Pam3Cys
(Fig. 3 Aand B) or Sendai virus (Fig. S7A) was largely preserved.
The defect in cytokine release could be explained by a decreased
NF-κB activation, as suggested by reduced IκBαdegradation,
IκBαphosphorylation, IKKα/βphosphorylation, and p65 nuclear
translocation, particularly at later time points during Y. pestis
or LPS challenge (Fig. 3 GJand Fig. S7 Band D). Reduced
signaling could also be observed in RIP1
/
macrophages (Fig. 3E
WT, uninfected WT, Y. pess RIP3- /-Casp8 -/-,Y.pess
C57Bl/6
RIP3-/-Casp8-/-
RIP3-/-Casp8+/-
RIP3-/-
0
20000
40000
60000
80000
100000 Medium
Y. pestis
BA
C
D
Salmonella
LPS+Nigericin
0
20
40
60
80
100
**
0
20
40
60
80 WT
RIP3-/-
RIP3-/-Casp8-/-
**
**
Y.pestis
Y.pseudo.
Caspase 8 activity (RLU)
*
*
Y.pestis
% cell death
Fig. 1. Caspase-8RIP3-deficient macrophages are protected against Y. pestis
induced cytotoxicity. (Aand B) Caspase-8
/
RIP3
/
(dKO), but not RIP3 KO
BMDM, are protected from Yersinia-induced cytotoxicity measured by LDH
release assay or (C) electron microscopy. (Scale bars, 2 μm.) Asterisks in Cin-
dicate bacteria. (D) Caspase-8 activity induced by Y. pestis infection (MOI 40,
2 h) in WT, RIP3
/
and dKO BMDMs. BMDM were infected with 1040 MOI of
Yersiniae or 1.5 MOI of Salmonella typhimurium for 4 h (Aand B)or2h(C
and D), and gentamycin was added after 1 h. Figures are representative for
three to eight experiments performed. Bars indicate mean plus SD. **P<0.01
(two-tailed ttest).
7392
|
www.pnas.org/cgi/doi/10.1073/pnas.1403477111 Weng et al.
and Fig. S7E). How caspase-8 controls NF-κB activation is unclear
and may not involve the enzymatic activity of caspase-8 (39) (Fig.
S7C); However, TRIF-mediated pathways may be targeted be-
cause MyD88-dependent TLR2 signaling is not affected.
Subsequent experiments indicated that YopJ-dependent
Y. pestis-induced IL-1βor IL-18 release was reduced in the ab-
sence of caspase-8 and showed further reduction by the absence
or blockade of RIP3 or RIP1 kinase activity (Fig. 4 AE).
B
A
Y.pestis
Salmonella
0
50
100
150
% c ell d eath
RIP1+/+
RI P1-/ -
**
**
RIP1+/+
RIP1-/-
0
20000
40000
60000
80000 Medium
Y.pestis
Caspase 8 activity (R LU)
No Ab
Medium
Y.pestis
YopJ
IP, RIP1 Ab
Caspase-8
RIP1
*
D
C
RIP1 INH GSK’963
CTR GSK’962
RIP3 INH GSK’872
-
-
-
+
-
-
-
+
-
-
-
+
Y.pestis
0
20
40
60
80 WT
% c el l deat h
Caspase 8 CKO
**
**
0
20000
40000
60000
80000
Caspase 8activity (RLU)
RIP1 INH GSK’963 --
+
Y.pestis
**
EF
Medium
Y.pestis
Medium
Y
.pestis
Medium
Y.pestis
Medium
Y.pestis
WT DKO RIP3-/- Casp8 CKO
RIP1
RIP3
β-actin
-70KD
-35KD
Fig. 2. RIP1 inhibition or deficiency protect macrophages from Y. pestis-induced cell death. (A) RIP1-deficient fetal liver macrophages are resistant to Y. pestis-
induced killing (MOI 40, 4 h), detected by LDH release. (Band D) RIP1, but not RIP3, mediates caspase-8 enzymatic activity after infection of BMDM (D) or fetal
liver macrophages (B) with Y. pestis for 2 h. (C) Caspase-8 conditional KO macrophages are protected from Y. pestis-induced death in the presence of RIP1
(GSK963) or RIP3 (GSK872) kinase inhibitors, but not by inactive compound GSK962. (D) RIP1 kinase inhibitor GSK963 inhibits caspase-8 enzyme activity after
infection. (E) RIP1 forms a complex with caspase-8 upon infection (1 h), measured by co-IP. (F) RIP1 is cleaved after Y. pestis infection in a caspase-8 dependent
fashion. Figures are representative for three to eight experiments performed. Bars indicate mean plus SD. **P<0.01 (two-tailed ttest).
ns
***
*
Medium
0
2
4
6
8
10
WT
Caspase 8 CKO
*
**
**
**
**
ΔΔ Δ
WT RIP3-/-Casp8-/- RIP3-/-
pro IL-1β-
β-actin-
Un Y.pe s t i s YopJ Un Y.p e s ti s YopJ Un Y. p e s ti s YopJ
0 15 30 45 0 15 30 45
(min)
WT RIP3-/-Casp8-/-
p-IĸBα
IĸBα
β-actin
Y.pestis ΔYopJ
0 15 30 45 0 15 30 45
WT RIP3-/-Casp8-/-
LPS (min)
pIKKα/β
IKKα
β-actin
B
EF
G
H
I
J
AC
D
IκBα
β-actin
WT RIP3-/-Casp8-/- RIP3-/-
0 15 30 45 60 80 0 15 30 45 60 80
WT RIP3-/-Casp8-/-
p-IκBα
IκBα
β-actin
LPS (min)
**
*
Y. pestis
Medium
Pam3Cys
LPS
Y.pestis
Salmonella
0
5
10
15
20 RIP1+/+
RIP1-/-
IL-6 (ng/mL)
Medium
Y.pestis
LPS
0
5
10
15
WT
RIP3-/-Casp8-/-
IL-1βmRNA
(Arbitrary units)
ns
**
Medium
0
1
2
3
4WT
RIP3-/-
RIP3-/- Casp8-/-
Y. pestis
Pam3Cys
0
2
4
6
8WT
RIP3-/-
RIP3-/- Casp8-/-
Y. pestis
Pam3Cys
Medium
LPS
0
5
10
15
IL-6 (ng/mL)
Medium
0
1
2
3
4
TNFα(ng/mL)
IL-6 (ng/mL)
Medium
Y. pestis
Y. pseudo.
LPS
Fig. 3. Caspase-8 and RIP1 contribute to cytokine
release and NF-κB activation. (AC) WT or mutant
BMDM were infected with Y. pestis,Y. pseudotu-
berculosis (MOI 10), or Salmonella (Sal, MOI 1.5) or
treated with LPS (100 ng/mL) or Pam3Cys (500 ng/
mL) for 6 h, and cytokine release was measured by
ELISA. (D) BMDMs were infected with Y. pestis for
4 h, mRNA was isolated, and quantitative PCR for
pro-IL-1βwas performed. (E) WT or RIP1
/
fetal
liver macrophages were stimulated with LPS (50
ng/mL), Pam3Cys (500 ng/mL), Y. pestis (MOI 10),
or Salmonella (MOI 1.5) for 6 h. IL-6 release was
measured by ELISA. (F) BMDMs were infected for
6 h and cell lysates probed for pro-IL-1β.(GJ)
BMDMs were infected or treated with LPS, mouse
TNF-α(10 min), and cell lysates were probed by
immunoblot for the indicated proteins (IκBα,
phospho-IκBα, phospho-IKKα/β,orβ-actin). Figures
are representative of two to five experiments
performed. Bars indicate means plus SD. **P<
0.01, *P<0.05 (two-tailed ttest).
Weng et al. PNAS
|
May 20, 2014
|
vol. 111
|
no. 20
|
7393
IMMUNOLOGY
Cytokine release after stimulation with Pam3Cys and nigericin
was unaffected (Fig. 4 Aand B), implying that NLRP3 activation
was not decreased. Although the absence of caspase-8 alone in
macrophages decreased IL-1βrelease induced by infection, it
increased IL-1βinduced by LPS alone (Fig. 4E), as suggested for
dendritic cells (40). Thus, more complex stimulations, as ob-
served during infection, yield a different result than a purified
ligand, possibly reflecting combined effects induced by both LPS
and the Yersinia T3SS in the context of live bacteria.
RIP1, Caspase-8, and RIP3 Mediate Inflammasome Activation. Our
previous data (Fig. 3) could partially explain reduced IL-1β
release. However, caspase-8 or RIP1 inhibition minimized
infection-induced caspase-1 cleavage (Fig. S7 Fand G), in-
dicating direct effects on inflammasome action. RIP3 has
been involved in inflammasome activation under certain con-
ditions with cIAP inhibiton (41). Infection-induced caspase-1 pro-
cessing, only partially dependent upon NLRP12 (25), was not
affected in RIP3
/
cells but was reduced in TLR4 or TRIF KO
(Fig. S8 Aand B) and caspase-8 cKO cells, and severely reduced in
RIP1
/
or RIP3
/
caspase-8
/
cells after Y. pestis infection (Fig. 5
A,B,andD). IL-1βprocessing was also affected (Fig. 5 Cand D).
Caspase-1 cleavage induced by Salmonella and Pam3Cys plus
nigericin was not affected (Fig. 5 Aand B), indicating that NLRC4-
and NLRP3-mediated caspase-1 cleavage is not inherently reduced
in RIP1
/
or dKO cells. Caspase-8 has been proposed to control
IL-1βmaturation and release in response to FasL stimulation or
fungal and bacterial challenge (42, 43), perhaps by directly cleaving
pro-IL-1β(44), and we cannot exclude this possibility. However, we
propose that caspase-8 directs caspase-1 processing and activation,
in a RIP3-enhanced manner, after Y. pestis challenge (Fig. 5 Band
D), but caspase-1 does not control caspase-8 activation (Fig. S8C).
Caspase-8 may be a critical component, but deletion or inhibition of
RIP3 may block an alternative pathway in the absence of caspase-8,
redundancy between caspase-8 and RIP3 may occur, or both mol-
ecules may be needed for stabilization of a signaling complex. The
mechanism we describe seems independent of FasL, TNF, or type
IIFN(Fig. S8 DF) and may have some common features with
responses induced by ER stress, certain chemotherapeutic drugs, or
Citrobacter (4547). The effect on caspase-1 cleavage may be me-
diated by the inflammasome adaptor ASC (Fig. 5E), because ASC
can associate with caspase-8 after Francisella or Salmonella in-
fection (38, 48), although the role of ASC may differ depending
upon conditions and source of YopJ (16, 23, 25).
Role of Caspase-8 and RIP3 for in Vivo Resistance to Bacterial
Infection. The in vivo relevance of our findings was emphasized
by the fact that RIP3
/
caspase-8
/
mice were more susceptible
to s.c. infection with virulent Y. pestis KIM1001 (Fig. 6A). Be-
cause LD
50
is very low for KIM1001 we used the attenuated
strain KIM1001-EcLpxL, which constitutively generates a TLR4-
activating hexa-acylated LPS (24), for survival analysis. dKO, or
lethally irradiated WT mice with bone marrow transplant (BMT)
from dKO, succumbed to s.c. infection with Y. pestis-EcLpxL
(Fig. 6Band Fig. S9A). Resistance to Y. pestis-LpxL is heavily
influenced by IL-18 and IL-1 (25). Moribund mice had large
numbers of bacteria in their spleens compared with WT controls,
suggesting that death occurred from uncontrolled systemic bac-
terial replication (Fig. 6C). This correlated with depressed IL-18,
IL-1β, TNF, and IL-6 cytokine levels and reduced myeloid cell
death (cells positive for live/dead stain and annexin V) in spleens
(Fig. 6 DIand Fig. S9 BG) after i.v. infection. Reduced ability
to suppress bacterial growth was also suggested by the presence
of visible bacteria-containing pockets in inflammatory foci in the
livers (Fig. 6J) of dKO BMT mice upon i.v. infection. Because
irradiated mice that received RIP3
/
caspase-8
/
BMT behaved
similarly as dKO animals, we propose that protection toward
infection is mediated by cells originating from the bone marrow,
expressing caspase-8 and RIP3. Some questions still remain with
respect to certain details of how caspase-8 and RIP3 are involved
in caspase-1 processing, although it is possible that ASC has
a central role. Our results provide a basis for increased un-
derstanding of how bacterial pathogens, via their T3SS, can in-
teract with several aspects of host innate immunity via RIP
kinases and caspase-8. The data also show how apoptosis, gen-
erally viewed as a silentcell death, can be accompanied by
strong inflammatory reactions, via pathways with several com-
mon players. The host may have developed these pathways as an
effective means of alerting cells to the infection. We propose that
caspase-8 and RIP kinases are central regulators of cell death
and innate immune responses to Y. pestis, and we establish a role
for these components in antibacterial innate immune responses.
Therapies that modulate the activity of these pathways may be
useful in the treatment of bacterial infections.
Methods
Mice. RIP3 KO (49) and caspase-8
/
RIP3
/
(dKO) (9) have been reported.
Caspase-8
fl/fl
LysM cre
+/+
cKO mice were generated by D.M.S. C57BL/6 mice
were bred in house or from Jackson Laboratories. BMT was performed on
lethally (900 rads) irradiated mice. Mice were infected s.c. or i.v. with 500 cfu
of KIM1001-pEcLpxL and monitored for survival. Tissue for analysis was
harvested at 42 h after infection, or at 68 h after s.c. infection with KIM1001
(300 cfu).
Bacterial Strains and Growth Conditions. Y. pestis KIM5 or KIM5ΔYopJ (24)
(25) were grown in tryptose-beef extract broth with 2.5 mM CaCl
2
overnight
with shaking at 26 °C. The next day the bacteria was diluted 1:8 in fresh
media, cultured for 1 h at 26 °C, and shifted to 37 °C for 2 h or grown
continuously at 37 °C when indicated. Y. pseudotuberculosis IP2666,
Y. enterocolitica 8081, and Salmonella enterica serovar Typhimurium strain
SL1344 were as reported (25) and grown at 37 °C. KIM5-EcLpxL and
KIM1001-EcLpxL were as previously published (24).
**
Medium
Y. pestis
Y. pestis∆YopJ
0.0
0.5
1.0
1.5
WT
RIP1 INH GSK’963 ---+
RIP3 INH GSK’872 --+-
Y.pestis
**
**
**
***
LPS
0
200
400
600
800
1000 WT
Caspase 8 CKO
** **
**
B
A
CDE
WT
RIP3-/-
RIP3-/-Casp8-/-
IL-1β (pg/mL)
0.0
0.2
0.4
0.6
0.8
1.0
5
10
15
WT
Caspase 8 CKO
**
Medium
Y. pestis
Pam3+Nigericin
IL-1β (ng/mL)
Caspase 8 CKO
IL-1β (ng/mL)
IL-18 (pg/mL)
Medium
Y. pestis
Y. pestis∆YopJ
IL-1β (pg/mL)
Medium
Y. pestis
LPS+Nec-1
***
ns
Medium
Pam3+Nigericin
IL-1β (ng/mL)
0.0
0.5
1.0
1.5
2.0
2.5
4
6
8
10
Y. pestis
0
100
200
300
400
0
50
100
150
200 WT
RIP3-/-
RIP3-/-Casp8-/-
Fig. 4. Y. pestis-induced release of IL-1βand IL-18 is severely reduced in
caspase-8/RIP3-deficient macrophages. (AE) BMDMs were infected with
Y. pestis or Y. pestis ΔYopJ for 6 h as indicated in Fig. 1, or stimulated with
nigericin (10 μg/mL) for 1 h after priming with Pam3Cys (4 h, 500 ng/mL).
IL-1βand IL-18 were analyzed by ELISA. (C) Some BMDMs were treated with
RIP1 inhibitor GSK963 (1 μM) or RIP3 inhibitor GSK872 (10 μM) for 1 h be-
fore infection. (E) BMDMs were challenged with Y. pestis (MOI 10) for 6 h or
LPS (50 ng/mL) for 10 h with or without Nec-1 pretreatment (20 μM). Figures
are representative of three to five experiments. Bars indicate means plus SD.
**P<0.01, *P<0.05 (two-tailed ttest in A,B, and D, and two-way ANOVA
with Tukeys posttest in Cand E).
7394
|
www.pnas.org/cgi/doi/10.1073/pnas.1403477111 Weng et al.
Cell Stimulations. BMDMs were prepared by maturing bone marrow cells for6
7 d in the presence of L929 supernatant containing M-CSF. Some experiments
were performed with BMDM immortalized with J2 retrovirus (42), or J2 im-
mortalized RIP1
+/+
and RIP1
/
fetal liver macrophages (31). Cells were plated
overnight and infected with bacteria at multiplicities of infection (MOIs) of 10
or 40, or stimulated with LPS from Y. pestis 26 °C (24) or Escherichia coli,or
Pam3Cys (Invivogen). Gentamycin was added 12 h after infection. Cell death
wasestimatedat4hbymeasuringlactate dehydrogenase (LDH) release
(Promega). In some experiments, cells were pretreated with 1 μMGSK963
or GSK962, or 3 μMGSK872 [RIP1 and RIP3 inhibitors (32) and GSK: P.A.H.,
J.B., P.J.G.], 20 μM Nec-1 (Enzo), 20 μM zIETD, zYVAD, or zVAD (Promega)
for 1 h before infection. Cytokines and caspase-1 cleavage were measured
Medium Y. p e s ti s Y.e n t e r c o
.Sal
.
Pam3+Nigericin
Pro Caspase-1
Caspase-1 p20
Lysate
Pro Caspase-1
SN
WT DKO RIP3-/- WT DKO RIP3-/- WT DKO RIP3-/-
Y. p e st i s S a l Pam3+Nigericin
Pro Caspase-1
Caspase-1 p20
SN
Lysate
WT DKO RIP3-/-
Medium
Pro Caspase-1
β-actin
WT DKO RIP3-/-
WT DKO RIP3-/- WT DKO RIP3-/-
Medium Y.pestis Pam3+Nigericin
-IL-1β p17
-pro IL-1β
SN
Lysate
-β-actin
Medium
Y.pestis
Pam3+Y.pestis
Medium
Y
.pestis
Pam3+Y.pestis
WT Caspase 8 CKO
Pro IL-1β
Pro IL-1β
IL-1β p17
Pro Caspase-1
Pro Caspase-1
Caspase-1 p20
β-actin
Medium
Y.pestis
Pam3+Niger.
Medium
Y.pestis
Pam3+Niger.
WT ASC-/-
Pro Caspase-1
Caspase-1 p20
AB
C
D
E
Fig. 5. RIP kinases and caspase-8 control caspase-1 cleavage induced by Y. pestis.(AE) BMDM (WT, RIP3
/
, RIP3
/
caspase-8
/
dKO, caspase-8 cKO) or fetal
liver macrophages (RIP1
+/+
, RIP1
/
) were infected with Y. pestis,Y. enterocolitica,orSalmonella (Sal) for 6 h or primed with Pam3Cys followed by nigericin for
1 h, and supernatants (SN) or lysates were analyzed for caspase-1 or IL-1βprocessing by immunoblots. Figures are representative of three to five experiments.
WT RIP3-/-Casp8-/-
WT
RIP3-/-Casp8-/-
1
2
3
4
5
6
7
8
9
10
**
Detection limit
WT RIP3-/-Casp8-/-
Annexin V
Live/Dead Blue
Y.pestis infected
2.05% 25.4%
9.56%62.9%
11.8%0.656%
7.35%80.2%
WT
RIP3-/-Casp8-/-
0
10
20
30
40
% of CD11b +Cel ls
Annexin V+, Live/Dead Blue+
Uninfected
WT
0
20
40
60
80
IL-1β(ng/g)
Uninfected
WT
RIP3-/-Casp8-/-
0.00
0.75
1.50
40
80
IL-6 (ng/g)
**
AB D
EFG I
H
C
J
WT
2
3
4
5
6
7
8
9
10
RIP3-/-Casp8-/-
CFU/spleen (log)
*
*
***
*
CFU/spleen (log)
RIP3-/-Casp8-/-
0
200
400
600
800
TNFa (pg/g)
Uninfected
WT
RIP3-/-Casp8-/-
0.0
0.5
1.0
25
50
IL-18 ( ng/g)
**
Uninfected WT
Uninfected DKO
WT
RIP3-/-Casp8-/-
p<0.001
Day
0 5 10 15 20
0
20
40
60
80
100
C57Bl/6 (n=12)
RIP3-/-Casp8-/- (n=12)
RIP3-/-Casp8+/- (n=12)
RIP3-/- (n=12)
Percent Servival
Fig. 6. Caspase-8 with RIP3 is critical for in
vivo resistance to bacterial infection. RIP3
/
caspase-8
/
dKO or WT mice were infected
s.c. with virulent Y. pestis KIM1001 (300 cfu)
for 68 h and spleens analyzed for bacterial
growth (A). Lethally irradiated mice, sub-
jected to bone marrow transplantation (BMT)
from the indicated genotypes (Band C), were
infected s.c. with 500 cfu of Y. pestis
KIM1001-EcLpxL and monitored for survival
(B), P<0.001 dKO vs. WT (logrank test).
Spleens from moribund dKO BMT mice and
controls were analyzed for bacterial contents
(C). (DJ) Mice from BMT as above (DF,HJ)
or regular dKO (G) were infected i.v. with
KIM1001-EcLpxL (500 cfu) for 42 h. Spleens
were homogenized and analyzed for cyto-
kines by ELISA (as cytokine/g tissue) (DG). (H
and I) CD11b-positive myeloid cells in spleens
were analyzed for cell death with live/dead
blue and annexin V stain. (J) Liver sections
were stainedwith hematoxylin and eosin and
subjected to microscopy (400×). Foci con-
taining inflammatory cells (mostly neu-
trophils) are shown, with visible pockets
containing bacteria indicated by arrows.
Shown is a representative experiment out of
two to three performed. *P<0.05, **P<0.01
(Mann-Whitney Utest).
Weng et al. PNAS
|
May 20, 2014
|
vol. 111
|
no. 20
|
7395
IMMUNOLOGY
as previously indicated (25). Caspase-8 activity (Promega) was measured
after 2 h.
ACKNOWLEDGMENTS. We thank Kelly Army, Gail Germain, and Anna
Cerny for help with mice; Shubhendu Ghosh for assistance with the
manuscript; TeChen Tzeng for help with microscopy; Vishva Dixit
(Genentech, Inc.) for providing RIP3 KO; Douglas R. Green and Christo-
pher Dillon for sending caspase-8 RIP3 dKO mice; Joan Mecsas and Mary
ORiordan for providing Y. pseudotuberculosis,Y. enterocolitica,and
Salmonella;D.M.S.(dmitryshay@emory.edu) for sharing cells from
previously unpublished casp8 cKO mice; and GSK: P.A.H., J.B., P.J.G. (peter.j.
gough@gsk.com), for providing RIP3 inhibitors and previously unpublished
RIP1 inhibitors. The work was supported by National Institutes of Health
(NIH) Grants AI07538 and AI057588-American Recovery and Reinvestment
Act (to E.L.), AI060025 (to N.S.), AI64349 and AI083713 (to K.A.F.), and
AI095213 (to G.I.V. and N.S.), the Norwegian Cancer Society, and the Research
Council of Norway. The study also used core services supported by University of
Massachusetts Diabetes and Endocrinology Research Center Gr ant D K325 20 and
the University of Massachusetts Core Electron MicroscopyFacility (supported
by NIH/National Center for Research Resources Award S10RR027897).
1. Lilo S, Zheng Y, Bliska JB (2008) Caspase-1 activation in macrophages infected with
Yersinia pestis KIM requires the type III secretion system effector YopJ. Infect Immun
76(9):39113923.
2. Philip NH, Brodsky IE (2012) Cell death programs in Yersinia immunity and patho-
genesis. Front Cell Infect Microbiol 2:149.
3. Lukaszewski RA, et al. (2005) Pathogenesis of Yersinia pestis infection in BALB/c mice:
Effects on host macrophages and neutrophils. Infect Immun 73(11):71427150.
4. Miao EA, et al. (2010) Caspase-1-induced pyroptosis is an innate immune effector
mechanism against intracellular bacteria. Nat Immunol 11(12):11361142.
5. Lamkanfi M, Dixit VM (2010) Manipulation of host cell death pathways during mi-
crobial infections. Cell Host Microbe 8(1):4454.
6. Vanlangenakker N, Vanden Berghe T, Vandenabeele P (2012) Many stimuli pull the
necrotic trigger, an overview. Cell Death Differ 19(1):7586.
7. Mocarski ES, Upton JW, Kaiser WJ (2012) Viral infection and the evolution of caspase
8-regulated apoptotic and necrotic death pathways. Nat Rev Immunol 12(2):7988.
8. Oberst A, et al. (2011) Catalytic activity of the caspase-8-FLIP(L) complex inhibits
RIPK3-dependent necrosis. Nature 471(7338):363367.
9. Kaiser WJ, et al. (2011) RIP3 mediates the embryonic lethality of caspase-8-deficient
mice. Nature 471(7338):368372.
10. Su H, et al. (2005) Requirement for caspase-8 in NF-kappaB activation by antigen
receptor. Science 307(5714):14651468.
11. Gröbner S, et al. (2007) Catalytically active Yersinia outer protein P induces cleavage
of RIP and caspase-8 at the level of the DISC independently of death receptors in
dendritic cells. Apoptosis 12(10):18131825.
12. Zheng Y, Lilo S, Mena P, Bliska JB (2012) YopJ-induced caspase-1 activation in Yersi-
nia-infected macrophages: Independent of apoptosis, linked to necrosis, dispensable
for innate host defense. PLoS ONE 7(4):e36019.
13. Monack DM, Mecsas J, Bouley D, Falkow S (1998) Yersinia-induced apoptosis in vivo
aids in the establishment of a systemic infection of mice. J Exp Med 188(11):
21272137.
14. Zhang Y, Bliska JB (2010) YopJ-promoted cytotoxicity and systemic colonization are
associated with high levels of murine interleukin-18, gamma interferon, and neu-
trophils in a live vaccine model of Yersinia pseudotuberculosis infection. Infect Immun
78(5):23292341.
15. Brodsky IE, Medzhitov R (2008) Reduced secretion of YopJ by Yersinia limits in vivo
cell death but enhances bacterial virulence. PLoS Pathog 4(5):e1000067.
16. Brodsky IE, et al. (2010) A Yersinia effector protein promotes virulence by preventing
inflammasome recognition of the type III secretion system. Cell Host Microbe 7(5):
376387.
17. Mukherjee S, et al. (2006) Yersinia YopJ acetylates and inhibits kinase activation by
blocking phosphorylation. Science 312(5777):12111214.
18. Mittal R, Peak-Chew SY, McMahon HT (2006) Acetylation of MEK2 and I kappa B
kinase (IKK) activation loop residues by YopJ inhibits signaling. Proc Natl Acad Sci USA
103(49):1857418579.
19. Paquette N, et al. (2012) Serine/threonine acetylation of TGFβ-activated kinase (TAK1)
by Yersinia pestis YopJ inhibits innate immune signaling. Proc Natl Acad Sci USA
109(31):1271012715.
20. Meinzer U, et al. (2012) Yersinia pseudotuberculosis effector YopJ subverts the Nod2/
RICK/TAK1 pathway and activates caspase-1 to induce intestinal barrier dysfunction.
Cell Host Microbe 11(4):337351.
21. Greten FR, et al. (2007) NF-kappaB is a negative regulator of IL-1beta secretion as
revealed by genetic and pharmacological inhibition of IKKbeta. Cell 130(5):918931.
22. Vladimer GI, Marty-Roix R, Ghosh S, Weng D, Lien E (2013) Inflammasomes and host
defenses against bacterial infections. Curr Opin Microbiol 16(1):2331.
23. Zheng Y, etal. (2011) A Yersinia effectorwith enhanced inhibitory activity on theNF-κB
pathway activates the NLRP3/ASC/caspase-1 inflammasome in macrophages. PLoS Pathog
7(4):e1002026.
24. Montminy SW, et al. (2006) Virulence factors of Yersinia pestis are overcome by a
strong lipopolysaccharide response. Nat Immunol 7(10):10661073.
25. Vladimer GI, et al. (2012) The NLRP12 inflammasome recognizes Yersinia pestis.Im-
munity 37(1):96107.
26. Oberst A, Green DR (2011) It cuts both ways: Reconciling the dual roles of caspase 8 in
cell death and survival. Nat Rev Mol Cell Biol 12(11):757763.
27. Cho YS, et al. (2009) Phosphorylation-driven assembly of the RIP1-RIP3 complex reg-
ulates programmed necrosis and virus-induced inflammation. Cell 137(6):11121123.
28. Zhang DW, et al. (2009) RIP3, an energy metabolism regulator that switches TNF-
induced cell death from apoptosis to necrosis. Science 325(5938):332336.
29. Moquin D, Chan FK (2010) The molecular regulation of programmed necrotic cell
injury. Trends Biochem Sci 35(8):434441.
30. Cusson-Hermance N, Khurana S, Lee TH, Fitzgerald KA, Kelliher MA (2005) Rip1
mediates the Trif-dependent toll-like receptor 3- and 4-induced NF-kappaB activation
but does not contribute to interferon regulatory factor 3 activation. J Biol Chem
280(44):3656036566.
31. Kelliher MA, et al. (1998) The death domain kinase RIP mediates the TNF-induced
NF-kappaB signal. Immunity 8(3):297303.
32. Kaiser WJ, et al. (2013) Toll-like receptor 3-mediated necrosis via TRIF, RIP3, and
MLKL. J Biol Chem 288(43):3126831279.
33. Gröbner S, et al. (2007) Absence of Toll-like receptor 4 signaling results in delayed
Yersinia enterocolitica YopP-induced cell death of dendritic cells. Infect Immun 75(1):
512517.
34. Zhang Y, Bliska JB (2003) Role of Toll-like receptor signaling in the apoptotic response
of macrophages to Yersinia infection. Infect Immun 71(3):15131519.
35. Haase R, et al. (2003) A dominant role of Toll-like receptor 4 in the signaling of ap-
optosis in bacteria-faced macrophages. J Immunol 171(8):42944303.
36. Kang TB, et al. (2004) Caspase-8 serves both apoptotic and nonapoptotic roles.
J Immunol 173(5):29762984.
37. Feoktistova M, et al. (2011) cIAPs block Ripoptosome formation, a RIP1/caspase-8
containing intracellular cell death complex differentially regulated by cFLIP isoforms.
Mol Cell 43(3):449463.
38. Man SM, et al. (2013) Salmonella infection induces recruitment of Caspase-8 to the
inflammasome to modulate IL-1βproduction. J Immunol 191(10):52395246.
39. Kang TB, et al. (2008) Mutation of a self-processing site in caspase-8 compromises its
apoptotic but not its nonapoptotic functions in bacterial artificial chromosome-
transgenic mice. J Immunol 181(4):25222532.
40. Kang TB, Yang SH, Toth B, Kovalenko A, Wallach D (2013) Caspase-8 blocks kinase
RIPK3-mediated activation of the NLRP3 inflammasome. Immunity 38(1):2740.
41. Vince JE, et al. (2012) Inhibitor of apoptosis proteins limit RIP3 kinase-dependent
interleukin-1 activation. Immunity 36(2):215227.
42. Bossaller L, et al. (2012) Cutting edge: FAS (CD95) mediates noncanonical IL-1βand IL-
18 maturation via caspase-8 in an RIP3-independent manner. J Immunol 189(12):
55085512.
43. Gringhuis SI, et al. (2012) Dectin-1 is an extracellular pathogen sensor for the in-
duction and processing of IL-1βvia a noncanonical caspase-8 inflammasome. Nat
Immunol 13(3):246254.
44. Maelfait J, et al. (2008) Stimulation of Toll-like receptor 3 and 4 induces interleukin-
1beta maturation by caspase-8. J Exp Med 205(9):19671973.
45. Antonopoulos C, El Sanadi C, Kaiser WJ, Mocarski ES, Dubyak GR (2013) Proapoptotic
chemotherapeutic drugs induce noncanonical processing and release of IL-1βvia
caspase-8 in dendritic cells. J Immunol 191(9):47894803.
46. Gurung P, et al. (2014) FADD and caspase-8 mediate priming and activation of the
canonical and noncanonical Nlrp3 inflammasomes. J Immunol 192(4):18351846.
47. Shenderov K, et al. (2014) Cutting edge: Endoplasmic reticulum stress licenses mac-
rophages to produce mature IL-1βin response to TLR4 stimulation through a caspase-
8- and TRIF-dependent pathway. J Immunol 192(5):20292033.
48. Pierini R, et al. (2012) AIM2/ASC triggers caspase-8-dependent apoptosis in Franci-
sella-infected caspase-1-deficient macrophages. Cell Death Differ 19(10):17091721.
49. Newton K, Sun X, Dixit VM (2004) Kinase RIP3 is dispensable for normal NF-kappa Bs,
signaling by the B-cell and T-cell receptors, tumor necrosis factor receptor 1, and Toll-
like receptors 2 and 4. Mol Cell Biol 24(4):14641469.
7396
|
www.pnas.org/cgi/doi/10.1073/pnas.1403477111 Weng et al.
... Journal of Experimental Medicine adapter RIPK1 (Monack et al., 1997;Palmer et al., 1998;Orth et al., 1999;Yoon et al., 2003;Mukherjee et al., 2006;Philip et al., 2014;Peterson et al., 2017). We previously demonstrated that mice lacking RIPK1 kinase activity (Ripk1 K45A ) in hematopoietic cells fail to form intact MLN PG and rapidly succumb to Yp infection (Peterson et al., 2017). ...
... WT Yp (clinical isolate strain 32777, serogroup O1) (Simonet and Falkow, 1992) and isogenic YopJ-deficient mutant were provided by Dr. James Bliska (Dartmouth College, Hanover, NH, USA) and previously described (Philip et al., 2014). Mutants lacking YopE (ΔyopE), YopH enzymatic activity (YopH R409A ), or both (denoted yopEH) were previously described (Sorobetea et al., 2023). ...
Article
Full-text available
Tumor necrosis factor (TNF) is a pleiotropic inflammatory cytokine that mediates antimicrobial defense and granuloma formation in response to infection by numerous pathogens. We previously reported that Yersinia pseudotuberculosis colonizes the intestinal mucosa and induces the recruitment of neutrophils and inflammatory monocytes into organized immune structures termed pyogranulomas (PG) that control Yersinia infection. Inflammatory monocytes are essential for the control and clearance of Yersinia within intestinal PG, but how monocytes mediate Yersinia restriction is poorly understood. Here, we demonstrate that TNF signaling in monocytes is required for bacterial containment following enteric Yersinia infection. We further show that monocyte-intrinsic TNFR1 signaling drives the production of monocyte-derived interleukin-1 (IL-1), which signals through IL-1 receptors on non-hematopoietic cells to enable PG-mediated control of intestinal Yersinia infection. Altogether, our work reveals a monocyte-intrinsic TNF-IL-1 collaborative inflammatory circuit that restricts intestinal Yersinia infection.
... The tandem DED domains (tDEDs) are responsible for Casp8 recruitment 91 , with the domains binding to signalling molecules such as FADD in the induction of apoptosis 92 . The tDEDs are used in Casp8 recruitment for multiple other signalling pathways, including ones relating to inflammation 93,94 . This is interesting, as longevity and suppressed inflammation are notable traits in bats 95 . ...
Article
Full-text available
Cancer is a disease that many multicellular organisms have faced for millions of years, and species have evolved various tumour suppression mechanisms to control oncogenesis. Although cancer occurs across the tree of life, cancer related mortality risks vary across mammalian orders, with Carnivorans particularly affected. Evolutionary theory predicts different selection pressures on genes associated with cancer progression and suppression, including oncogenes, tumour suppressor genes and immune genes. Therefore, we investigated the evolutionary history of cancer associated gene sequences across 384 mammalian taxa, to detect signatures of selection across categories of oncogenes (GRB2, FGL2 and CDC42), tumour suppressors (LITAF, Casp8 and BRCA2) and immune genes (IL2, CD274 and B2M). This approach allowed us to conduct a fine scale analysis of gene wide and site-specific signatures of selection across mammalian lineages under the lens of cancer susceptibility. Phylogenetic analyses revealed that for most species the evolution of cancer associated genes follows the species’ evolution. The gene wide selection analyses revealed oncogenes being the most conserved, tumour suppressor and immune genes having similar amounts of episodic diversifying selection. Despite BRCA2’s status as a key caretaker gene, episodic diversifying selection was detected across mammals. The site-specific selection analyses revealed that the two apoptosis associated domains of the Casp8 gene of bats (Chiroptera) are under opposing forces of selection (positive and negative respectively), highlighting the importance of site-specific selection analyses to understand the evolution of highly complex gene families. Our results highlighted the need to critically assess different types of selection pressure on cancer associated genes when investigating evolutionary adaptations to cancer across the tree of life. This study provides an extensive assessment of cancer associated genes in mammals with highly representative, and substantially large sample size for a comparative genomic analysis in the field and identifies various avenues for future research into the mechanisms of cancer resistance and susceptibility in mammals.
... Besides its major role in cell apoptosis promoted by death receptors triggering, mitochondrial apoptosis, and endoplasmic reticulum stress [59][60][61], CASP8 has been shown to regulate inflammation by modulating IL-1β mRNA expression, specifically by the activation of nuclear factor-kB (NF-κB) [62]. In other biological scenarios, CASP8 has also been involved in NLRP3 priming and activation, where it was directly associated with cleavage and processing of procaspase-1, IL-1β, and IL-18 processing [63][64][65][66]. We decided to study CASP8 in our model due to its contribution to neuronal pathologies like TBI [67][68][69], brain ischemia [70], and seizures [71]. ...
Article
Full-text available
Traumatic brain injury is a leading cause of disability and death worldwide and represents a high economic burden for families and national health systems. After mechanical impact to the head, the first stage of the damage comprising edema, physical damage, and cell loss gives rise to a second phase characterized by glial activation, increased oxidative stress and excitotoxicity, mitochondrial damage, and exacerbated neuroinflammatory state, among other molecular calamities. Inflammation strongly influences the molecular events involved in the pathogenesis of TBI. Therefore, several components of the inflammatory cascade have been targeted in experimental therapies. Application of Electromagnetic Field (EMF) stimulation has been found to be effective in some inflammatory conditions. However, its effect in the neuronal recovery after TBI is not known. In this pilot study, Yucatan miniswine were subjected to TBI using controlled cortical impact approach. EMF stimulation via a helmet was applied immediately or two days after mechanical impact. Three weeks later, inflammatory markers were assessed in the brain tissues of injured and contralateral non-injured areas of control and EMF-treated animals by histomorphometry, immunohistochemistry, RT-qPCR, Western blot, and ELISA. Our results revealed that EMF stimulation induced beneficial effect with the preservation of neuronal tissue morphology as well as the reduction of inflammatory markers at the transcriptional and translational levels. Immediate EMF application showed better resolution of inflammation. Although further studies are warranted, our findings contribute to the notion that EMF stimulation could be an effective therapeutic approach in TBI patients.
Article
Gasdermin D (GSDMD) is a pore-forming protein that perforates the plasma membrane (PM) during pyroptosis, a pro-inflammatory form of cell death, to induce the unconventional secretion of inflammatory cytokines and, ultimately, cell lysis. GSDMD is activated by protease-mediated cleavage of its active N-terminal domain from the autoinhibitory C-terminal domain. Inflammatory caspase-1, -4/5 are the main activators of GSDMD via either the canonical or non-canonical pathways of inflammasome activation, but under certain stimuli, caspase-8 and other proteases can also activate GSDMD. Activated GSDMD can oligomerize and assemble into various nanostructures of different sizes and shapes that perforate cellular membranes, suggesting plasticity in pore formation. Although the exact mechanism of pore formation has not yet been deciphered, cysteine residues are emerging as crucial modulators of the oligomerization process. GSDMD pores and thus the outcome of pyroptosis can be modulated by various regulatory mechanisms. These include availability of activated GSDMD at the PM, control of the number of GSDMD pores by PM repair mechanisms, modulation of the lipid environment and post-translational modifications. Here, we review the latest findings on the mechanisms that induce GSDMD to form membrane pores and how they can be tightly regulated for cell content release and cell fate modulation.
Preprint
Full-text available
Infection with the helminth Schistosoma mansoni can cause exacerbated morbidity and mortality via a pathogenic host CD4 T cell-mediated immune response directed against parasite egg antigens, with T helper (Th) 17 cells playing a major role in the development of severe granulomatous hepatic immunopathology. The role of inflammasomes in intensifying disease has been reported; however, neither the types of caspases and inflammasomes involved, nor their impact on the Th17 response are known. Here we show that enhanced egg-induced IL-1β secretion and pyroptotic cell death required both caspase-1 and caspase-8 as well as NLRP3 and AIM2 inflammasome activation. Schistosome genomic DNA activated AIM2, whereas reactive oxygen species, potassium efflux and cathepsin B, were the major activators of NLRP3. NLRP3 and AIM2 deficiency led to a significant reduction in pathogenic Th17 responses, suggesting their crucial and non-redundant role in promoting inflammation. Additionally, we show that NLRP3- and AIM2-induced IL-1β suppressed IL-4 and protective Type I IFN (IFN-I) production, which further enhanced inflammation. IFN-I signaling also curbed inflammasome-mediated IL-1β production suggesting that these two antagonistic pathways shape the severity of disease. Lastly, Gasdermin D (Gsdmd) deficiency resulted in a marked decrease in egg-induced granulomatous inflammation. Our findings establish NLRP3/AIM2-Gsdmd axis as a central inducer of pathogenic Th17 responses which is counteracted by IFN-I pathway in schistosomiasis. Summary Schistosomiasis is a major tropical parasitic disease caused by trematode worms of the genus Schistosoma. Morbidity and mortality in infection with the species Schistosoma mansoni are due to a pathogenic CD4 T cell-mediated immune response directed against parasite eggs, resulting in granulomatous inflammation. In severe cases of schistosomiasis, there is liver fibrosis, hepatosplenomegaly, portal hypertension, gastro-intestinal hemorrhage and death. Here we describe the role of two proteins, the NLRP3 and AIM2 inflammasomes, in intensifying disease. We found that upstream proteins which activate these inflammasomes are caspase-1 and caspase 8; these in turn lead to the activation of another protein, Gasdermin D (Gsdmd), which facilitates the release of the proinflammatory cytokine IL-1β. Importantly, we observed that mice deficient in Gsdmd exhibit diminished pathology. Finally, we discovered that the protective Type I Interferon (IFN-I) pathway counteracts the caspase/inflammasome/Gsdmd axis thereby controlling egg mediated inflammation. These results give us a deeper understanding of the functional features of the crosstalk between inflammasome and IFN-I pathway, which may lead to the identification of novel targets for therapeutic intervention.
Preprint
Full-text available
Coxiella burnetii is an obligate intracellular bacteria which causes the global zoonotic disease Q Fever. Treatment options for infection are limited, and development of novel therapeutic strategies requires a greater understanding of how C. burnetii interacts with immune signaling. Cell death responses are known to be manipulated by C. burnetii , but the role of caspase-8, a central regulator of multiple cell death pathways, has not been investigated. In this research, we studied bacterial manipulation of caspase-8 signaling and the significance of caspase-8 to C. burnetii infection, examining bacterial replication, cell death induction, and cytokine signaling. We measured caspase, RIPK, and MLKL activation in C. burnetii -infected TNFα/CHX-treated THP-1 macrophage-like cells and TNFα/ZVAD-treated L929 cells to assess apoptosis and necroptosis signaling. Additionally, we measured C. burnetii replication, cell death, and TNFα induction over 12 days in RIPK1-kinase-dead, RIPK3-kinase-dead, or RIPK3-kinase-dead-caspase-8 -/- BMDMs to understand the significance of caspase-8 and RIPK1/3 during infection. We found that caspase-8 is inhibited by C. burnetii , coinciding with inhibition of apoptosis and increased susceptibility to necroptosis. Furthermore, C. burnetii replication was increased in BMDMs lacking caspase-8, but not in those lacking RIPK1/3 kinase activity, corresponding with decreased TNFα production and reduced cell death. As TNFα is associated with the control of C. burnetii , this lack of a TNFα response may allow for the unchecked bacterial growth we saw in caspase-8 -/- BMDMs. This research identifies and explores caspase-8 as a key regulator of C. burnetii infection, opening novel therapeutic doors.
Article
Inflammasome-mediated caspase-1 activation facilitates innate immune control of Plasmodium in the liver, thereby limiting the incidence and severity of clinical malaria. However, caspase-1 processing occurs incompletely in both mouse and human hepatocytes and precludes the generation of mature IL-1β or IL-18, unlike in other cells. Why this is so or how it impacts Plasmodium control in the liver has remained unknown. We show that an inherently reduced expression of the inflammasome adaptor molecule apoptosis-associated specklike protein containing CARD (ASC) is responsible for the incomplete proteolytic processing of caspase-1 in murine hepatocytes. Transgenically enhancing ASC expression in hepatocytes enabled complete caspase-1 processing, enhanced pyroptotic cell death, maturation of the proinflammatory cytokines IL-1β and IL-18 that was otherwise absent, and better overall control of Plasmodium infection in the liver of mice. This, however, impeded the protection offered by live attenuated antimalarial vaccination. Tempering ASC expression in mouse macrophages, on the other hand, resulted in incomplete processing of caspase-1. Our work shows how caspase-1 activation and function in host cells are fundamentally defined by ASC expression and offers a potential new pathway to create better disease and vaccination outcomes by modifying the latter.
Article
Full-text available
Programmed necrosis is a form of caspase-independent cell death whose molecular regulation is poorly understood. The kinase RIP1 is crucial for programmed necrosis, but also mediates activation of the prosurvival transcription factor NF-κB. We postulated that ...
Article
Full-text available
The Nlrp3 inflammasome is critical for host immunity, but the mechanisms controlling its activation are enigmatic. In this study, we show that loss of FADD or caspase-8 in a RIP3-deficient background, but not RIP3 deficiency alone, hampered transcriptional priming and posttranslational activation of the canonical and noncanonical Nlrp3 inflammasome. Deletion of caspase-8 in the presence or absence of RIP3 inhibited caspase-1 and caspase-11 activation by Nlrp3 stimuli but not the Nlrc4 inflammasome. In addition, FADD deletion prevented caspase-8 maturation, positioning FADD upstream of caspase-8. Consequently, FADD- and caspase-8-deficient mice had impaired IL-1β production when challenged with LPS or infected with the enteropathogen Citrobacter rodentium. Thus, our results reveal FADD and caspase-8 as apical mediators of canonical and noncanonical Nlrp3 inflammasome priming and activation.
Article
Full-text available
Caspase-8 is now appreciated to govern both apoptosis following death receptor ligation and cell survival and growth via inhibition of the Ripoptosome. Cells must therefore carefully regulate the high level of caspase-8 activity during apoptosis versus the modest levels observed during cell growth. The caspase-8 paralogue c-FLIP is a good candidate for a molecular rheostat of caspase-8 activity. c-FLIP can inhibit death receptor-mediated apoptosis by competing with caspase-8 for recruitment to FADD. However, full-length c-FLIPL can also heterodimerize with caspase-8 independent of death receptor ligation and activate caspase-8 via an activation loop in the C terminus of c-FLIPL. This triggers cleavage of c-FLIPL at Asp-376 by caspase-8 to produce p43FLIP. The continued function of p43FLIP has, however, not been determined. We demonstrate that acute deletion of endogenous c-FLIP in murine effector T cells results in loss of caspase-8 activity and cell death. The lethality and caspase-8 activity can both be rescued by the transgenic expression of p43FLIP. Furthermore, p43FLIP associates with Raf1, TRAF2, and RIPK1, which augments ERK and NF-κB activation, IL-2 production, and T cell proliferation. Thus, not only is c-FLIP the initiator of caspase-8 activity during T cell activation, it is also an initial caspase-8 substrate, with cleaved p43FLIP serving to both stabilize caspase-8 activity and promote activation of pathways involved with T cell growth.
Article
Full-text available
Nucleotide-binding oligomerization domain-like receptors (NLRs) detect pathogens and danger-associated signals within the cell. Salmonella enterica serovar Typhimurium, an intracellular pathogen, activates caspase-1 required for the processing of the proinflammatory cytokines, pro-IL-1β and pro-IL-18, and pyroptosis. In this study, we show that Salmonella infection induces the formation of an apoptosis-associated specklike protein containing a CARD (ASC)-Caspase-8-Caspase-1 inflammasome in macrophages. Caspase-8 and caspase-1 are recruited to the ASC focus independently of one other. Salmonella infection initiates caspase-8 proteolysis in a manner dependent on NLRC4 and ASC, but not NLRP3, caspase-1 or caspase-11. Caspase-8 primarily mediates the synthesis of pro-IL-1β, but is dispensable for Salmonella-induced cell death. Overall, our findings highlight that the ASC inflammasome can recruit different members of the caspase family to induce distinct effector functions in response to Salmonella infection.
Article
Full-text available
The identification of noncanonical (caspase-1-independent) pathways for IL-1β production has unveiled an intricate interplay between inflammatory and death-inducing signaling platforms. We found a heretofore unappreciated role for caspase-8 as a major pathway for IL-1β processing and release in murine bone marrow-derived dendritic cells (BMDC) costimulated with TLR4 agonists and proapoptotic chemotherapeutic agents such as doxorubicin (Dox) or staurosporine (STS). The ability of Dox to stimulate release of mature (17-kDa) IL-1β was nearly equivalent in wild-type (WT) BMDC, Casp1(-/-)Casp11(-/-) BMDC, WT BMDC treated with the caspase-1 inhibitor YVAD, and BMDC lacking the inflammasome regulators ASC, NLRP3, or NLRC4. Notably, Dox-induced production of mature IL-1β was temporally correlated with caspase-8 activation in WT cells and greatly suppressed in Casp8(-/-)Rip3(-/-) or Trif(-/-) BMDC, as well as in WT BMDC treated with the caspase-8 inhibitor, IETD. Similarly, STS stimulated robust IL-1β processing and release in Casp1(-/-)Casp11(-/-) BMDC that was IETD sensitive. These data suggest that TLR4 induces assembly of caspase-8-based signaling complexes that become licensed as IL-1β-converting enzymes in response to Dox and STS. The responses were temporally correlated with downregulation of cellular inhibitor of apoptosis protein 1, suggesting suppressive roles for this and likely other inhibitor of apoptosis proteins on the stability and/or proteolytic activity of the caspase-8 platforms. Thus, proapoptotic chemotherapeutic agents stimulate the caspase-8-mediated processing and release of IL-1β, implicating direct effects of such drugs on a noncanonical inflammatory cascade that may modulate immune responses in tumor microenvironments.
Article
Full-text available
Toll-like receptor (TLR) signaling is triggered by pathogen-associated molecular patterns that mediate well established cytokine-driven pathways, activating NF-κB together with IRF3/IRF7. In addition, TLR3 drives caspase 8-regulated programmed cell death pathways reminiscent of TNF family death receptor signaling. We find that inhibition or elimination of caspase 8 during stimulation of TLR2, TLR3, TLR4, TLR5, or TLR9 results in receptor interacting protein (RIP) 3 kinase-dependent programmed necrosis that occurs through either TIR domain-containing adapter-inducing interferon-β (TRIF) or MyD88 signal transduction. TLR3 or TLR4 directly activates programmed necrosis through a RIP homotypic interaction motif-dependent association of TRIF with RIP3 kinase (also called RIPK3). In fibroblasts, this pathway proceeds independent of RIP1 or its kinase activity, but it remains dependent on mixed lineage kinase domain-like protein (MLKL) downstream of RIP3 kinase. Here, we describe two small molecule RIP3 kinase inhibitors and employ them to demonstrate the common requirement for RIP3 kinase in programmed necrosis induced by RIP1-RIP3, DAI-RIP3, and TRIF-RIP3 complexes. Cell fate decisions following TLR signaling parallel death receptor signaling and rely on caspase 8 to suppress RIP3-dependent programmed necrosis whether initiated directly by a TRIF-RIP3-MLKL pathway or indirectly via TNF activation and the RIP1-RIP3-MLKL necroptosis pathway. Background: RIP3-dependent programmed necrosis is an alternative to apoptosis. Results: When caspase-8 is compromised, TRIF-dependent TLRs directly activate RIP3 kinase through RHIM-dependent interactions. Conclusion: TRIF mediates direct RHIM-dependent signaling, triggering necrosis via RIP3 and MLKL. Significance: Programmed necrosis eliminates cells following stimulation of either MyD88 or TRIF signaling pathways that converge on RIP3.
Article
Full-text available
Recognizing the presence of invading pathogens is key to mounting an effective innate immune response. Mammalian cells express different classes of germline-encoded pattern recognition receptors that monitor the extracellular and intracellular compartments of host cells for signs of infection and that activate several conserved signalling pathways. An efficient immune response often requires the sequential detection of a pathogen by different receptors in different subcellular compartments, which results in a complex interplay of downstream signalling pathways. In this Review, we discuss the recent identification of previously unknown pattern recognition receptors and how they complement the repertoire of established receptors.
Article
Full-text available
In addition to its pro-apoptotic function in the death receptor pathway, roles for caspase-8 in mediating T-cell proliferation, maintaining lymphocyte homeostasis, and suppressing immunodeficiency have become evident. Humans with a germline point mutation of CASPASE-8 have multiple defects in T cells, B cells, and NK cells, most notably attenuated activation and immunodeficiency. By generating mice with B-cell-specific inactivation of caspase-8 (bcasp8-/-), we show that caspase-8 is dispensable for B-cell development, but its loss in B cells results in attenuated antibody production upon in vivo viral infection. We also report an important role for caspase-8 in maintaining B-cell survival following stimulation of the Toll-like receptor (TLR)2, -3, and -4. In response to TLR4 stimulation, caspase-8 is recruited to a complex containing IKKαβ, and its loss resulted in delayed NFκB nuclear translocation and impaired NFκB transcriptional activity. Our study supports dual roles for caspase-8 in apoptotic and nonapoptotic functions and demonstrates its requirement for TLR signaling and in the regulation of NFκB function.
Article
The accumulation of improperly folded proteins within the endoplasmic reticulum (ER) generates perturbations known as ER stress that engage the unfolded protein response. ER stress is involved in many inflammatory pathologies that are also associated with the production of the proinflammatory cytokine IL-1β. In this study, we demonstrate that macrophages undergoing ER stress are able to drive the production and processing of pro-IL-1β in response to LPS stimulation in vitro. Interestingly, the classical NLRP3 inflammasome is dispensable, because maturation of pro-IL-1β occurs normally in the absence of the adaptor protein ASC. In contrast, processing of pro-IL-1β is fully dependent on caspase-8. Intriguingly, we found that neither the unfolded protein response transcription factors XBP1 and CHOP nor the TLR4 adaptor molecule MyD88 is necessary for caspase-8 activation. Instead, both caspase activation and IL-1β production require the alternative TLR4 adaptor TRIF. This pathway may contribute to IL-1-driven tissue pathology in certain disease settings.