Rana et al Final (ref comm edit) - Spiral

advertisement
Bacterial TIR containing proteins and host innate immune system evasion
Rohini R. Rana1+, Minghao Zhang1, Abigail M. Spear2, Helen S. Atkins2, Bernadette
Byrne1*
Division of Molecular Biosciences, Imperial College London, London SW7 2AZ,
UK1 and Department of Biomedical Sciences, Defence Science and Technology
Laboratory, Porton Down, Salisbury, Wiltshire, SP4 0JQ, UK2.
*Corresponding
Dr Bernadette Byrne,
Division of Molecular Biosciences
Imperial College London
South Kensington
London
SW7 2AZ.
Fax: +44 20 7594 3022
E-mail: b.byrne@imperial.ac.uk
+ Now based at MRC National Institute for Medical Research, The Ridgeway, Mill
Hill, London NW7 1AA, UK.
@ Crown Copyright 2011. Published with the permission of the Defence Science and
Technology Laboratory on behalf of the Controller of HMSO.
1
Abstract
The innate immune system provides the first line of host defence against
invading pathogens. Key to upregulation of the innate immune response are Toll-like
receptors (TLRs) which recognise pathogen-associated molecular patterns (PAMPs)
and trigger a signaling pathway culminating in the production of inflammatory
mediators. Central to this TLR signaling pathway are heterotypic protein-protein
interactions mediated through Toll/Interleukin-1 receptor (TIR) domains found in
both the cytoplasmic regions of TLRs and adaptor proteins. Pathogenic bacteria have
developed a range of ingenuous strategies to evade the host immune mechanisms.
Recent work has identified a potentially novel evasion mechanism involving bacterial
TIR domain proteins. Such domains have been identified in a wide range of
pathogenic bacteria and there is evidence to suggest that they interfere directly with
the TLR signaling pathway and thus inhibit the activation of NFB. The individual
TIR domains from the pathogenic bacteria Salmonella enterica serovar Enteritidis,
Brucella sp, uropathogenic E. coli and Yersinia pestis have been analysed in detail.
The individual bacterial TIR domains from these pathogenic bacteria seem to differ in
their modes of action and their roles in virulence. Here we review the current state of
knowledge on the possible roles and mechanisms of action of the bacterial TIR
domains.
Keywords: TLR signaling pathway, TIR domain, innate immunity, pathogenic
bacteria, bacterial TIR domain, virulence
2
The innate immune system and TLRs
Upregulation of the innate immune system provides the first line of host
defence against invading pathogens. Key components of the innate immune system
are pattern recognition receptors (PRRs) which recognise pathogen-associated
molecular patterns (PAMPs) and initiate an intracellular signaling cascade
culminating in a host immune response [1]. The best-studied PRRs are the Toll-like
receptors (TLRs), type I transmembrane sensory receptors that recognise a range of
PAMPs from pathogenic bacteria, fungi, viruses and parasitic-protozoa [1]. The
principal role of the TLRs is in regulation of the innate immune system however there
is evidence to show that they also play important functions in shaping adaptive
immune responses, particularly in combination with other PRRs [2].
TLRs are expressed on immune cells such as phagocytes, B cells and certain T
cells, and on non-immune cells including fibroblasts and epithelial cells. They can be
located on the cell surface (TLR1-2, TLR4-6 and TLR10) or in intracellular
compartments such as endosomes (TLR3 and TLR7-9) [1,3].
Ten TLRs are encoded in the human genome, each of which has evolved to
recognise certain PAMPs from a broad range of microbes. For instance, TLRs signal
the presence of triacyl lipopeptides (Pam3CSK4; TLR1/2), peptidoglycan and
zymosan
(TLR2),
double-stranded
ribonucleic
acid
(ds
RNA;
TLR3),
lipopolysaccharides (LPS) of Gram-negative bacteria (TLR4), flagellin (TLR5),
single-stranded RNA (ss RNA; TLR7 and TLR8) and unmethylated CpG DNA
(TLR9) [3,4]. The ligand for TLR10 is currently unknown. The expression of the
different TLRs is subject to change and is regulated by rapid cellular responses to
pathogens, various cytokines and environmental stresses [1].
3
The TLRs contain an extracellular leucine-rich repeat (LRR) domain that
recognises and binds to specific microbial ligands [5]. Linked to the LRR domain by a
single transmembrane domain [6] is a cytoplasmically located Toll/Interleukin-1
receptor (TIR) domain that mediates downstream signaling [7]. TLRs are thought to
exist as dimers, which can be heterotypic, homotypic or both depending of the
receptor [8,9]. Upon interaction with a PAMP the TLR dimer undergoes a molecular
rearrangement of the intracellular TIR domains to form an active interaction domain
allowing recruitment of adaptor molecules [4,8,9]. TLR signalling pathways converge
to activate transcription factors that regulate the induction of proinflammatory
cytokines and co-stimulatory molecules [5]. This in turn elicits innate immune
responses against the invading pathogens.
One of the key features of TLR signaling is that the signal is both initiated and
propagated by heterotypic TIR-TIR interactions (Fig 1). In addition to the TIR
domains of the TLR receptors there are five TIR containing adaptor proteins [10];
myeloid differentiation factor 88 (MyD88), TIR domain containing adaptor protein
including interferon- (TRIF), MyD88-adaptor like (MAL, also called TIRAP),
TRIF-related adaptor molecule (TRAM) and sterile - and armadillo-motifcontaining protein (SARM). The MyD88-mediated pathway leads to the activation of
NFκB and MAP kinase pathways inducing the production and release of
proinflammatory cytokines [5]. The TRIF-mediated pathway leads to activation of
interferon response factors (IRFs) inducing the production of type I interferon [11].
All TLRs and IL-1Rs, except TLR3, signal through the MyD88-dependent pathway
while TLR3 and TLR4 function through the MyD88-independent pathway (TRIF;
[12]). TLR1, TLR2, TLR4 and TLR6 additionally recruit the TIR domain bridging
adaptor MAL, which mediates the interactions between the TIR domains of TLRs and
4
MyD88 [13]. Similarly, TLR4 recruits the alternative TIR domain bridging adaptor,
TRAM, in addition to TRIF to link between the TIR domain of TLR4 and TRIF [14]
(Fig 1). The final TIR containing adaptor, SARM, acts as a negative regulator of
TRIF signaling [15].
TIR domains
The cytoplasmic TIR domains of TLRs typically contain 150-200 amino acid
residues [16]. As mentioned the principal role of TIR domains is the mediation of
heterotypic protein-protein interactions between receptor and adaptor molecules in the
signal transduction process. TIR domains share common sequence motifs, called box
1 (F/Y)-DAFISY), box 2 (GYKLC-RD-PG) and box 3 (a conserved W surrounded by
basic residues), of which boxes 1 and 2 are vital for mediating signalling [12,17] (Fig
2). One of the most functionally important regions of the TIR domains is the BB loop
(Fig 2 and Fig 3), part of the Box 2 region. This motif is comprised of the sequence
RDxPG (where x is any residue and  is a hydrophobic residue). Site-directed
mutagenesis studies have revealed that several residues within this loop are important
for TLR signal transduction. This is most notable in the case of the Pro residue.
Mutation of this residue, Pro712, to His in TLR4-TIR results in a loss of TLR
signalling in response to lipopolysaccharide [18].
A summary of all the current solved structures of TIR domains is given in
Table 1. The structure of the human TIR domain of TLR1 reveals a central fivestranded parallel β-sheet (βA - βE) surrounded by five helices (A - E) and the
connecting loops (Fig 2 and Fig 3) [19]. The functionally relevant BB loop (Fig 2 and
Fig 3) connects strand βB and helix B, which protrudes out from the main core of
the protein in most structures of TIR domains [16,19-22]. Whilst the core structure is
5
maintained in most TIR domains, some conformational differences are observed even
in the case of the TIR domains of TLR1 and TLR2 which share 50% sequence
identity [19]. The regions of the proteins that exhibit the largest variation include the
helices B and D and the loops BB, CD and DD. Such structural diversity is
suggested to be crucial for the specificity of signal transduction from different
receptors [19,23]. The recent structure of MAL lacks the helix connecting the B and
C strands. This region of the protein is present as a long flexible loop which links the
A helix and the B strand [22]. It is not clear precisely why there is less order in this
region of MAL compared to the other TIR domains however it seems likely that this
is important for function.
Microbial targeting of the TLR signaling pathway
Research over the last decade or so has highlighted that the TLR signaling pathway is
a key target of pathogen mechanisms of host immune system evasion. One of the
strategies for successful host evasion involves manipulation of the TLR signaling
pathway via modification of microbial PAMPs so that they are less effective at
stimulating the TLRs. An example of this can be seen with the modification in Lipid
A content of Yersinia pestis with changing growth temperature [24]. Alternatively, by
co-stimulating host carbohydrate binding receptors as well as TLRs it has been shown
that Mycobacterium tuberculosis [25] is able to initiate anti-inflammatory
mechanisms counteracting the effects of TLR signaling. Viruses have been shown to
target various levels of the TLR signaling pathway. The A46R and A52R proteins
from vaccinia virus reduce TLR signaling by forming interactions with a range of host
TIR containing proteins including MyD88, MAL, TRIF and TRAM [26] as well as
6
TLR4 [27]. Based on sequence homology to the Box2 region of TIR domains it was
thought that A46R and A52R function as negative regulators of the heterotypic TIRTIR interactions involved in upregulating the TLR signaling pathway through
competitive binding with host TIR domain proteins. However A52R has also been
shown to form direct association with Interleukin-1 receptor-associated kinase-like 2
(IRAK2) [28], a component of the TLR signaling pathway which does not contain a
TIR domain, so it may be that the role of this protein is more general. The suggestion
that the viral proteins are not TIR containing proteins, is supported by analysis using
bioinformatics tools which were unable to find known TIR domains using A46 as a
search protein or find TIR domains using A46 as a search protein [29]. There are
proteins in other viruses such as Hepatitis B that also interact with components of the
TLR signaling pathways probably in a similar way to the vaccinia proteins [30]. In
contrast, the Karposi’s sarcoma virus replication and transcription activator (RTA)
causes degradation of TRIF [31].
Bacterial TIRs and their prevalence throughout bacterial species
A new potential host evasion mechanism involving TLRs came to light with the
identification of bacterial TIR homologues. The first report described the
identification >200 TIR homologues in a wide range of bacterial species including
Staphylococcus aureus, Brucella melitensis and Salmonella enterica serovar
Enteriditis [32]. Sequence alignment of the so-called TIR-like protein A (TlpA) from
S. enterica serovar Enteriditis with the TIR domains of human TLR6, TLR4 and
MyD88, revealed high sequence homology in the box 1 region [32]. Subsequent work
identified Paracoccus denitrificans TIR-like protein (PdTLP) as a homologue of S.
7
enterica serovar Enteriditis tlpA, sharing 39% sequence similarity over 255 residues
[33]. Cirl et al. [34] identified bacterial TIR domains in the CFT073 strain of
uropathogenic Escherichia coli and in Brucella species, which were termed TIR
domain containing proteins C (TcpC) and B (TcpB, also known as Btp1). A more
comprehensive bioinformatics analysis by Spear et al. [29] identified 922 bacterial
TIR domains in a wide range of both pathogenic and non-pathogenic bacterial species
including the plague-causing bacterium, Yersinia pestis. More recently a study
identified a substantially lower number of 483 bacterial TIR domains proteins [35]. It
is possible that this is due to subtle differences in the search methodologies used to
find the proteins. An intriguing finding was revealed by recent research by Zhang et al
[35] demonstrating that animal SARM is more closely related to the bacterial TIR
domain proteins than it is to the other animal TIR adaptors suggesting a common
origin for these proteins.
The bacterial TIR proteins are approximately 230 to 310 amino acids long
with the conserved TIR domain comprised of 150-200 amino acids. The TIR domain
can be located in either the C-terminal or the N-terminal region and the remaining
region of the protein can be highly variable. In the TlpA and PdTLP proteins the TIR
domain is located within the C-terminal region with an N-terminal region postulated
to comprise a highly -helical coiled coil domain. The presence of a coiled coil
domain was supported by circular dichroism analysis on the N-terminal region of
PdTLP indicating a high -helical content [33]. In contrast in the case of TcpC the Nterminal region has been annotated as a putative transmembrane segment while the Nterminal domain of TcpB has been identified as a lipid binding domain with
specificity for phosphoinositide phosphates. The recent bioinformatic analysis
highlights the co-occurrence of the TIR domains with a range of different domains
8
[29]. Examples include the presence of an N-terminal TIR domain followed by Cterminal tetratricopeptide repeats (TPR), common protein-protein interaction
domains, in Frankia sp. [29]. Bacterial TIR containing proteins in Beggiatos sp.
contain an N-terminal Mettallophos domain involved in phosphorylation, while in
Clostridium thermocellum it contains C-terminal trypsin, a serine-protease domain,
and in Mariprofundus ferroxydans it contains a C-terminal Mrr-cat, a type IV
restriction endonuclease domain [29].
Bacterial TIR proteins and the TLR signaling pathway
The majority of studies on bacterial TIR proteins have focused on their
potential role as virulence factors that directly subvert host TLR signaling. Newman
et al. [32] in the first studies on a bacterial TIR showed that S. enterica serovar
Enteritidis TlpA suppresses the induction of NFB activation by the mammalian TIRcontaining proteins TLR4, IL-1R and MyD88 but does not suppress NFB activation
induced by the cytokine TNF in cultured mammalian cells. Most notably, mice
infected with genetic knockouts of S. enterica serovar Enteritidis lacking the TlpA
encoding gene showed increased time to death compared to mice infected with the
wild-type strain. The TlpA knockout strain also demonstrated a reduced ability to
survive in cultured mammalian cells [32]. These data strongly suggested that the TlpA
was important for virulence of S. enterica serovar Enteritidis via a mechanism of
inhibition of the TLR signaling pathway responsible for activation of NFB.
Support for both a role in virulence and the general mechanism of action of
bacterial TIR proteins came from further studies on homologues from Brucella
melitensis (TcpB [also known as Btp1]; [34,36]) and the uropathogenic E. coli strain
CFT073 (TcpC; [34]). One study revealed that both TcpB and TcpC suppress TLR2-
9
and TLR4-mediated activation of NFB but do not suppress the activation of NFB
induced by the TIR-independent cytokine TNF and that a TcpC knockout strain
showed a reduced ability to survive in mouse RAW264.7 macrophage cells compared
to wild-type CFT073 [34]. It was possible to achieve wild-type accumulation rates in
RAW264.7 cells when the TcpC knockout was complemented with a plasmid borne
copy of TcpC. In addition, a comparison using a mouse model of urinary tract
infections revealed that the TcpC knockout strain had reduced virulence compared to
wild-type [34]. A further in-depth analysis of the precise sites of activity of the
bacterial TIR proteins revealed that TcpB and the purified TIR domain of TcpC
(TcpC-TIR) interacted with endogenous MyD88 in mammalian cell lines [34]. Pulldown assays confirmed the interaction of TcpC-TIR directly with MyD88 but
revealed a lack of interaction with two other TIR containing mammalian proteins
TLR2-TIR and TRIF, or two non-TIR domain containing components of the TLR
signaling pathway, IRAK1 and IRAK4 [34]. More recently there is some evidence to
suggest that TcpC also interacts with TLR4 [37]. Other TIR containing proteins
involved in TLR signaling, for example MAL and TRAM, have not been tested so far
for interaction with TcpC.
Based on sequence similarity, plasma membrane localization profile and the
ability to inhibit MAL-induced NFkB activation it has been suggested that TcpB acts
as a MAL mimic and competes with MAL for binding of MyD88 [36]. However more
recent research suggests that TcpB only interacts with MAL and not MyD88 and does
not block the interaction between these two proteins [38]. Instead, data indicate that
TcpB interaction results in a reduction in the cellular levels of MAL in mammalian
cells expressing TcpB [38]. It has been suggested that interaction with TcpB results in
ubiquitination and targeting of MAL for degradation. Since TcpB is unlikely to act
10
directly as a ubiquitin ligase it is possible that TcpB binding causes degradation by
indirect means [38]. Studies on TcpB from Brucella abortus have identified a role in
inhibition of maturation of infected dendritic cells, suggested to act as key cellular
environments for pathogen proliferation [39].
Most recently research has focused on the TIR protein from the plague
causing bacterium, Yersinia pestis (YpTdp). In an in vitro reporter assay YpTdp
inhibits activation of NFB, although the TIR domain only (YpTIR) has no effect
[40]. However, pull-down assays show that YpTIR does form interactions with
MyD88-TIR [40] suggesting that in this case the N-terminal, non-TIR, domain of the
YpTdp also plays a role in inhibition of the TLR signaling pathway. Both inhibition
of NFκB by YpTdp and interaction of YpTIR with MyD88-TIR are dependent upon
the proline residue located on the BB loop [40]. This indicates that YpTdp mediates
the downregulation of the innate immune system through competitive interaction with
MyD88 as described for TcpC [34].
Studies with a genetic deletion mutant of Y. pestis indicate that YpTdp does
not have a role in virulence of the organism. It is possible that any effects on virulence
in this case are minor. Y. pestis employs a very sophisticated system of infection and
evasion mechanisms. For example YopJ, a protease delivered into host cells via the
type III section system (TTSS) is known to block NFB activation by TLR receptors
[41]. Since it appears that in Y. pestis there is some redundancy in the mechanisms
employed to inhibit NFB activation it is possible that the YopJ is masking the effects
of YpTdp on the virulence of Y. pestis.
The data reported so far on bacterial TIR proteins strongly indicates that there
is no one clear mechanism by which they act to subvert the host immune system.
11
TcpC and TlpA show direct interactions with components of the TLR signaling
pathway and are likely to directly block the pathway whereas TcpB seems to trigger
the degradation of one key component and thus reduce signaling. While it seems clear
that in some cases the bacterial TIR domains play a role in virulence in the case of
YpTdp, the protein lacks any clear relevance for virulence despite inhibiting NFB
activation. It seems likely that the TIR proteins from these pathogenic bacteria have
different roles which means that only by studying a range of these can we really
understand the full breadth of functions.
TIR domains in non-pathogenic bacteria
As mentioned above TIR domains have been identified in a range of both pathogenic
and non-pathogenic bacterial species [29]. To account for this it has been suggested
that the TIR domain is a generic protein-protein interaction domain which has
subsequently been adapted to function as a mechanism of host immune system
evasion [29]. However, whilst the roles of the TIR domains in the non-pathogenic
bacteria remain uncharacterized there is evidence to show that these can be used to
understand the mechanism of action of the TIR domains from pathogenic bacteria.
This is illustrated by research carried out on the TIR domain protein from Paracoccus
denitrificans (PdTdp). Intriguingly this protein interacts with both MyD88 and TLR4
[33] and is the only bacterial TIR protein for which there is a high resolution structure
([42]; see below for a more detailed discussion). This has allowed the first insights
into the domains involved in bacterial TIR domain interaction with MyD88.
Anecdotal reports and our own experiences indicate that bacterial TIR domain
proteins are challenging to express and purify. However the data from the P.
denitrificans studies indicates that PdTdp is a suitable homologue for studying some
aspects of the function of TIR domains in pathogenic bacteria.
12
How do the bacterial TIR proteins localise to the host cell cytoplasm
Sequence analysis of the regions around tlpA revealed a phage origin for tlpA
as well as links with pilin-related domains [32]. Since pilins are extracellular proteins
playing roles in cell adhesion, it was suggested that proteins encoded within this
genomic region including TlpA are secreted [32]. However further links with the type
III or type IV secretion systems could not be made on the basis of gene location due
to discontinuous genome sequence of S. enterica in this region [32]. To date, no more
direct evidence of secretion of TlpA has been reported.
TcpC has been reported to be secreted into the media of cultured bacterial
cells and taken up into host macrophages in a cholesterol-dependent manner to
interfere with TLR-mediated TNF induction [34]. The secretion mechanism has yet
to be identified for TcpC as the protein does not contain a recognizable signal
sequence. As mentioned previously, TcpB has been shown to contain an N-terminal
lipid binding domain which is suggested to important for cell permeability. Indeed
purified MBP-TcpB internalizes into mouse macrophages in a dose dependent manner
[43]. The YpTdp does not contain any known signal sequences and no mechanism of
secretion has yet been identified [40].
As yet there is no clear consensus on the mechanism by which TIR domains
localize to the host cytosol and the data so far indicates that there may be different
mechanisms depending on the pathogen. Further studies are needed to clarify this
issue perhaps by performing real-time imaging and co-localisation studies.
13
Biophysical characterization of bacterial TIR proteins
Preliminary biophysical characterization of bacterial TIR proteins by
molecular modelling indicated that they share considerable structural homology with
the TIR domain of TLR1 [34]. This was confirmed by the first crystal structure of a
bacterial TIR protein, PdTIR (Table 1), which revealed an overall fold identical to the
known structures of the human TIR domains [42]. The asymmetric unit of the PdTIR
crystals contains four molecules that form two equivalent dimers [42]. Unlike the
homodimer interface observed in the structure of the human TLR10-TIR [20]
involving mainly the BB loops (Figs 2 and 3; Table 1), the dimer interface in PdTIR
involves the D-helix, the DD- and EE-loops leaving the BB-loops largely exposed
on the surface of the molecules [42]. A large network of hydrogen bonds mediates
dimerisation of the PdTIR protomers in the crystal structure.
The crystal structures of a number of TIR domains [16,19,20,22,42] indicates
the formation of dimers in the crystal lattice. In contrast, the purified proteins are all
present as monomers in solution although there is some evidence to suggest that TIR
domains function as weakly associated dimers [44].
In contrast to all the other biochemically characterized TIR domains YpTIR
forms dimers in solution which are likely to be disulphide bond dependent [45].
Mutation of one of the two Cys residues to Ser present in YpTIR retains the dimeric
form of the protein whereas mutation of both results in highly unstable protein not
suitable for analysis. Interestingly this study showed that mutation of Cys132 to Ser
abolished interaction with MyD88 suggesting that the disulphide bonds may be key in
maintaining an active confirmation of YpTIR [45].
14
The structure of PdTIR indicated that the highly exposed BB loop is the region
of PdTIR which interacts with human TIR proteins of the TLR signaling pathway
[42]. The importance of the BB loop region in interactions between bacterial TIR
domains and human adaptor proteins has been demonstrated by mutagenesis studies.
Pull-down studies on a Pro173His mutant of YpTIR shows no detectable interaction
with GST-MyD88-TIR in contrast to the wild-type protein [40]. In addition, mutation
of another highly conserved residue of the BB loop Gly158 to an Ala resulted in a
TcpB with reduced activity in a NFkB reporter assay [36]. However it should be noted
that there is variability in this region of the bacterial TIR domains with TcpB and
YpTdp lacking the conserved Pro and Gly residues respectively. While these studies
strongly suggest that the BB loop is essential for bacterial TIR activity only a
structure of a bacterial TIR domain in complex with a human adaptor protein will
provide definitive characterization of the interaction domain.
Other roles of bacterial TIR domains
While the majority of studies have focused on the role of bacterial TIR
domains in direct inhibition of the TLR signaling pathway through competitive
binding of human adaptor proteins recent research has highlighted another possible
function. Radhakrishnan and colleagues [46] have shown that exposure of
microtubules to TcpB expression results in improved nucleation and growth of
microtubules in vitro. Preliminary results show that this activity is mediated by the
BB loop. The significance of this effect of TcpB in terms of immune system evasion
remains to be clarified although the authors of the study speculate that this may be a
15
further mechanism by which the upregulation of the NFkB signaling pathway is
inhibited.
The first study on bacterial TIR proteins indicated that expression of TlpA
from S. enterica serovar Enteritidis resulted in activation of caspase-1, a host protease
[32]. Activation of this protease results in cleavage, activation and secretion of the
proinflammatory cytokine IL-1. An increase in IL-1 secretion was observed in
cultured cells expressing TlpA and in macrophages infected with wild-type S.
enterica serovar Enteritidis compared to the tlpA deletion strain.
This rather
surprising finding suggests that TlpA is capable of both simulating and inhibiting
proinflammatory cytokine production. Recent research has indicated that YopJ from
Y. pestis also has such a dual role [47]. It has been suggested that the upregulation of
caspase-1 may represent a mechanism of host defence against pathogens able to
inhibit the NKB pathway. It will be interesting to see if bacterial TIR domain
proteins in addition to TlpA are able to stimulate this host cell response.
The YpTdp deletion mutant shows some interesting phenotypes including
increased auto-aggregation and reduced ability to survive in high salt conditions when
compared to wild-type Y. pestis [40]. These features may contribute to the ability of
the Y. pestis to survive in the host and thus increase pathogenicity. They also suggest
functions of the YpTdp in addition to inhibition of TLR signaling.
Overall, the widespread distribution of TIR domains in non-pathogenic
bacterial species as well as the diversity of the associated protein domains [29] is
strongly indicative that these TIR containing proteins have roles other than in immune
system evasion. However it remains to be seen what these other possible roles are.
16
Microbial TIR domains as therapeutic agents
Upregulation of the innate immune response is highly advantageous to host
survival although anomalous upregulation has been implicated in a number of
autoimmune and inflammatory diseases including multiple sclerosis and rheumatoid
arthritis [48]. In addition, research has also highlighted the role of NFB in cancer.
Overactivation of NFB as the result of mutagenesis of proteins in the TLR signaling
pathway may contribute to cell proliferation [49]. Recently, highly potent TLR
signaling inhibitors have been generated based on the A46R protein from vaccinia
virus [50]. Previous studies have shown it is possible to make cell-penetrating
peptides based on the BB-loops of the TIR domains [51] of both adaptor proteins and
TLRs which inhibit TLR signaling pathways. The wide variety of ways in which
bacterial TIR domains modulate the TLR signaling pathway means that it may be
possible to design peptides which target different components of the pathway and thus
have subtly different effects on the innate immune system. These represent a means of
tuning the innate immune response to possibly alleviate some of the symptoms of
debilitating auto-immune diseases and potentially treat some cancers.
Conclusion
The current evidence strongly suggests that bacterial TIR domains represent a novel
mechanism by which pathogens can subvert the host immune system. However
questions remain regarding how the bacterial TIR domains are taken up into host cells
and at what points in the TLR signaling pathway they can form negatively regulating
interactions. Some of the studies seem to indicate that while the bacterial TIR
domains share some common features there is a certain amount of individuality in
17
their interaction profiles and precise functions. Further research is required in order to
elucidate the precise mechanism of action of these interesting molecules.
Acknowledgements
This work was funded by the UK Ministry of Defence.
References
1.
Akira S, Uematsu S, Takeuchi O (2006) Pathogen recognition and innate
immunity. Cell 124:783–801.
2.
Fritz JH, Le Bourhis L, Sellge G, Magalhaes JG, Fsihi H, Kufer TA, Collins C,
Viala J, Ferrero RL, Girardin SE, et al. (2007) Nod1-mediated innate immune
recognition of peptidoglycan contributes to the onset of adaptive immunity.
Immunity 26:445–459.
3.
Kumar H, Kawai T, Akira S (2009) Pathogen recognition in the innate immune
response. Biochem J 420:1–16.
4.
Gay NJ, Gangloff M (2007) Structure and function of Toll receptors and their
ligands. Annu Rev Biochem 76:141–165.
5.
Akira S, Takeda K (2004) Toll-like receptor signalling. Nat Rev Immunol
4:499–511.
6.
Hashimoto C, Hudson KL, Anderson KV (1988) The Toll gene of Drosophila,
required for dorsal-ventral embryonic polarity, appears to encode a
transmembrane protein. Cell 52:269–279.
7.
Gay NJ, Keith FJ (1991) Drosophila Toll and IL-1 receptor. Nature 351:355–
356.
8.
Triantafilou M, Gamper FGJ, Haston RM, Mouratis MA, Morath S, Hartung T,
Triantafilou K (2006) Membrane sorting of toll-like receptor (TLR)-2/6 and
TLR2/1 heterodimers at the cell surface determines heterotypic associations
with CD36 and intracellular targeting. J Biol Chem 281:31002–31011.
9.
Latz E, Verma A, Visintin A, Gong M, Sirois CM, Klein DCG, Monks BG,
18
McKnight CJ, Lamphier MS, Duprex WP, et al. (2007) Ligand-induced
conformational changes allosterically activate Toll-like receptor 9. Nat
Immunol 8:772–779.
10.
O'Neill LAJ, Bowie AG (2007) The family of five: TIR-domain-containing
adaptors in Toll-like receptor signalling. Nat Rev Immunol 7:353–364.
11.
Núñez Miguel R, Wong J, Westoll JF, Brooks HJ, O'Neill LAJ, Gay NJ,
Bryant CE, Monie TP (2007) A dimer of the Toll-like receptor 4 cytoplasmic
domain provides a specific scaffold for the recruitment of signalling adaptor
proteins. PLoS ONE 2:e788.
12.
Brown V, Brown RA, Ozinsky A, Hesselberth JR, Fields S (2006) Binding
specificity of Toll-like receptor cytoplasmic domains. Eur. J. Immunol.
36:742–753.
13.
Horng T, Barton GM, Flavell RA, Medzhitov R (2002) The adaptor molecule
TIRAP provides signalling specificity for Toll-like receptors. Nature 420:329–
333.
14.
Rowe DC, McGettrick AF, Latz E, Monks BG, Gay NJ, Yamamoto M, Akira
S, O'Neill LA, Fitzgerald KA, Golenbock DT (2006) The myristoylation of
TRIF-related adaptor molecule is essential for Toll-like receptor 4 signal
transduction. Proc Natl Acad Sci USA 103:6299–6304.
15.
Carty M, Goodbody R, Schröder M, Stack J, Moynagh PN, Bowie AG (2006)
The human adaptor SARM negatively regulates adaptor protein TRIFdependent Toll-like receptor signaling. Nat Immunol 7:1074–1081.
16.
Khan JA, Brint EK, O'Neill LAJ, Tong L (2004) Crystal structure of the
Toll/interleukin-1 receptor domain of human IL-1RAPL. J Biol Chem
279:31664–31670.
17.
Slack JL, Schooley K, Bonnert TP, Mitcham JL, Qwarnstrom EE, Sims JE,
Dower SK (2000) Identification of two major sites in the type I interleukin-1
receptor cytoplasmic region responsible for coupling to pro-inflammatory
signaling pathways. J Biol Chem 275:4670–4678.
18.
Poltorak A, He X, Smirnova I, Liu MY, van Huffel C, Du X, Birdwell D,
Alejos E, Silva M, Galanos C, et al. (1998) Defective LPS signaling in
C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282:2085–
2088.
19.
Xu Y, Tao X, Shen B, Horng T, Medzhitov R, Manley JL, Tong L (2000)
Structural basis for signal transduction by the Toll/interleukin-1 receptor
domains. Nature 408:111–115.
20.
Nyman T, Stenmark P, Flodin S, Johansson I, Hammarström M, Nordlund P
(2008) The crystal structure of the human toll-like receptor 10 cytoplasmic
domain reveals a putative signaling dimer. J Biol Chem 283:11861–11865.
21.
Ohnishi H, Tochio H, Kato Z, Orii KE, Li A, Kimura T, Hiroaki H, Kondo N,
19
Shirakawa M (2009) Structural basis for the multiple interactions of the
MyD88 TIR domain in TLR4 signaling. Proc Natl Acad Sci USA 106:10260–
10265.
22.
Valkov E, Stamp A, Dimaio F, Baker D, Verstak B, Roversi P, Kellie S, Sweet
MJ, Mansell A, Gay NJ, et al. (2011) Crystal structure of Toll-like receptor
adaptor MAL/TIRAP reveals the molecular basis for signal transduction and
disease protection. Proc Natl Acad Sci USA 108:14879–14884.
23.
Watters TM, Kenny EF, O'Neill LAJ (2007) Structure, function and regulation
of the Toll/IL-1 receptor adaptor proteins. Immunol. Cell Biol. 85:411–419.
24.
Kawahara K, Tsukano H, Watanabe H, Lindner B, Matsuura M (2002)
Modification of the structure and activity of lipid A in Yersinia pestis
lipopolysaccharide by growth temperature. Infect Immun 70:4092–4098.
25.
Nigou J, Zelle-Rieser C, Gilleron M, Thurnher M, Puzo G (2001)
Mannosylated lipoarabinomannans inhibit IL-12 production by human
dendritic cells: evidence for a negative signal delivered through the mannose
receptor. J Immunol 166:7477–7485.
26.
Stack J, Haga IR, Schröder M, Bartlett NW, Maloney G, Reading PC,
Fitzgerald KA, Smith GL, Bowie AG (2005) Vaccinia virus protein A46R
targets multiple Toll-like-interleukin-1 receptor adaptors and contributes to
virulence. J Exp Med 201:1007–1018.
27.
Bowie A, Kiss-Toth E, Symons JA, Smith GL, Dower SK, O'Neill LA (2000)
A46R and A52R from vaccinia virus are antagonists of host IL-1 and toll-like
receptor signaling. Proc Natl Acad Sci USA 97:10162–10167.
28.
Harte MT, Haga IR, Maloney G, Gray P, Reading PC, Bartlett NW, Smith GL,
Bowie A, O'Neill LAJ (2003) The poxvirus protein A52R targets Toll-like
receptor signaling complexes to suppress host defense. J Exp Med 197:343–
351.
29.
Spear AM, Loman NJ, Atkins HS, Pallen MJ (2009) Microbial TIR domains:
not necessarily agents of subversion? Trends Microbiol 17:393–398.
30.
Lang T, Lo C, Skinner N, Locarnini S, Visvanathan K, Mansell A (2011) The
Hepatitis B e antigen (HBeAg) targets and suppresses activation of the Tolllike receptor signaling pathway. J Hepatol doi:10.1016/j.jhep.2010.12.042.
31.
Ahmad H, Gubbels R, Ehlers E, Meyer F, Waterbury T, Lin R, Zhang L (2011)
Kaposi sarcoma-associated herpesvirus degrades cellular Toll-interleukin-1
receptor domain-containing adaptor-inducing beta-interferon (TRIF). J Biol
Chem 286:7865–7872.
32.
Newman RM, Salunkhe P, Godzik A, Reed JC (2006) Identification and
characterization of a novel bacterial virulence factor that shares homology with
mammalian Toll/interleukin-1 receptor family proteins. Infect Immun 74:594–
601.
20
33.
Low LY, Mukasa T, Reed JC, Pascual J (2007) Characterization of a TIR-like
protein from Paracoccus denitrificans. Biochem Biophys Res Comm 356:481–
486.
34.
Cirl C, Wieser A, Yadav M, Duerr S, Schubert S, Fischer H, Stappert D,
Wantia N, Rodriguez N, Wagner H, et al. (2008) Subversion of Toll-like
receptor signaling by a unique family of bacterial Toll/interleukin-1 receptor
domain-containing proteins. Nat Med 14:399–406.
35.
Zhang Q, Zmasek CM, Cai X, Godzik A (2011) TIR domain-containing
adaptor SARM is a late addition to the ongoing microbe–host dialog. Develop
Comp Immunol 35:461–468.
36.
Radhakrishnan GK, Yu Q, Harms JS, Splitter GA (2009) Brucella TIR
Domain-containing Protein Mimics Properties of the Toll-like Receptor
Adaptor Protein TIRAP. J Biol Chem 284:9892–9898.
37.
Yadav M, Zhang J, Fischer H, Huang W, Lutay N, Cirl C, Lum J, Miethke T,
Svanborg C (2010) Inhibition of TIR domain signaling by TcpC: MyD88dependent and independent effects on Escherichia coli virulence. PLoS Pathog
6:e1001120.
38.
Sengupta D, Koblansky A, Gaines J, Brown T, West AP, Zhang D, Nishikawa
T, Park S-G, Roop RM, Ghosh S (2010) Subversion of innate immune
responses by Brucella through the targeted degradation of the TLR signaling
adapter, MAL. J Immunol 184:956–964.
39.
Salcedo SP, Marchesini MI, Lelouard H, Fugier E, Jolly G, Balor S, Muller A,
Lapaque N, Demaria O, Alexopoulou L, et al. (2008) Brucella control of
dendritic cell maturation is dependent on the TIR-containing protein Btp1.
PLoS Pathog 4:e21.
40.
Spear AM, Rana R, Jenner DC, Flick-Smith HC, Oyston PCF, Simpson P,
Matthews S, Byrne B, Atkins HS (2012) A TIR domain protein from Yersinia
pestis interacts with mammalian IL-1/TLR pathways but does not play a central
role in the virulence of Y. pestis in a mouse model of bubonic plague.
Microbiology 158:1593-606.
41.
Zhang Y, Ting AT, Marcu KB, Bliska JB (2005) Inhibition of MAPK and NFkappa B pathways is necessary for rapid apoptosis in macrophages infected
with Yersinia. J Immunol 174:7939–7949.
42.
Chan SL, Low LY, Hsu S, Li S, Liu T, Santelli E, Le Negrate G, Reed JC,
Woods VL, Pascual J (2009) Molecular mimicry in innate immunity: crystal
structure of a bacterial TIR domain. J Biol Chem 284:21386–21392.
43.
Radhakrishnan GK, Splitter GA (2010) Biochemical and functional analysis of
TIR domain containing protein from Brucella melitensis. Biochem Biophys
Res Comm 397:59–63.
44.
Dunne A, Ejdeback M, Ludidi PL, O'Neill LAJ, Gay NJ (2003) Structural
complementarity of Toll/interleukin-1 receptor domains in Toll-like receptors
21
and the adaptors Mal and MyD88. J Biol Chem 278:41443–41451.
45.
Rana RR, Simpson P, Zhang M, Jennions M, Ukegbu C, Spear AM, Alguel Y,
Matthews SJ, Atkins HS, Byrne B (2011) Yersinia pestis TIR-domain protein
forms dimers that interact with the human adaptor protein MyD88. Microb
Pathog 51:89–95.
46.
Radhakrishnan G, Harms J, Splitter G (2011) Modulation of microtubule
dynamics by a TIR domain protein from the intracellular pathogen, Brucella
melitensis. Biochem J 439:79-83.
47.
Zheng Y, Lilo S, Brodsky IE, Zhang Y, Medzhitov R, Marcu KB, Bliska JB
(2011) A Yersinia effector with enhanced inhibitory activity on the NF-κB
pathway activates the NLRP3/ASC/caspase-1 inflammasome in macrophages.
PLoS Pathog 7:e1002026.
48.
Beutler B (2004) Inferences, questions and possibilities in Toll-like receptor
signalling. Nature 430:257–263.
49.
Ngo VN, Young RM, Schmitz R, Jhavar S, Xiao W, Lim K-H, Kohlhammer H,
Xu W, Yang Y, Zhao H, et al. (2011) Oncogenically active MYD88 mutations
in human lymphoma. Nature 470:115–119.
50.
Lysakova-Devine T, Keogh B, Harrington B, Nagpal K, Halle A, Golenbock
DT, Monie T, Bowie AG (2010) Viral inhibitory peptide of TLR4, a peptide
derived from vaccinia protein A46, specifically inhibits TLR4 by directly
targeting MyD88 adaptor-like and TRIF-related adaptor molecule. J Immunol
185:4261–4271.
51.
Toshchakov VY, Fenton MJ, Vogel SN (2007) Cutting Edge: Differential
inhibition of TLR signaling pathways by cell-permeable peptides representing
BB loops of TLRs. J Immunol 178:2655–2660.
52.
Chan SL, Mukasa T, Santelli E, Low LY, Pascual J. (2010). The crystal
structure of a TIR domain from Arabidopsis thaliana reveals a conserved
helical region unique to plants. Protein Science, 19:155–161.
53.
Bernoux M, Ve T, Williams S, Warren C, Hatters D, Valkov E, Zhang, X., et
al. (2011). Structural and functional analysis of a plant resistance protein TIR
domain reveals interfaces for self-association, signaling, and autoregulation.
Cell Host Microbe 9:200–211.
Figure Legends
22
Figure 1 Overview of TLR/ IL-1R signalling through TIR domain containing
adaptors The receptors IL-1R, TLR1, TLR2, TLR4, TLR5 and TLR6 are expressed
on the cell surface while receptors TLR3, TLR7, TLR8 and TLR9 are expressed on
endosomal membranes. The cytoplasmic TIR domain containing adaptors MyD88,
Mal, TRIF and TRAM are differentially recruited by TLR/ IL-1R complexes to
positively regulate the activation of transcription factors NF-κB and IRFs while
SARM negatively regulates TRIF. The evidence so far indicates that bacterial TIR
proteins can also act as negative regulators through interaction with MyD88, MAL
and the TIR domain of TLR4 (indicated by the white asterisks; see main text for more
details). However the interaction profile differs for individual bacterial TIR domains.
Figure 2 Primary sequence alignment of a range of bacterial TIR domains with some
mammalian TIR domains of known structure. The regions corresponding to the TIR
motifs, Box1, 2 and 3, are indicated by the solid rectangles above the alignment. The
regions of the sequence of TLR1-TIR corresponding to secondary structure elements
as determined by X-ray crystallographic analysis (Pdb accession code: 1FYV) are
shown below the alignment. Some of the residues involved in dimer formation are
shown in bold for both the TIR domain of TLR10 and PdTIR. The alignment was
generated using ClustalW.
Figure 3 Structure of a) the TIR domain of TLR1 (Pdb accession code, 1FYV) and b)
PdTIR (Pdb accession code, 3H16). The positions of the BB (red), DD (green) and EE
(cyan) loops with key roles in dimer formation are indicated. c) Superposition of the
TIR domain of TLR1 (dark blue) and PdTIR (magenta). Space-filling model of the
structure of the dimer of d) TIR domain of TLR10 (Pdb accession code, 2J67) and e)
PdTIR. In each case monomer 1 is shown in dark blue with the region forming the
23
interaction in red while monomer 2 is shown in cyan with the regions forming the
interaction in gold. Some of the key residues involved in dimer formation are
individually labeled. In d) the residues indicated all form part of the BB loop while in
e) the residues are part of the DD and EE loops as indicated in a) and b). See also
Figure 2 for more details.
Figure 1
24
Figure 2
25
Figure 3
26
27
Download