首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Toll/IL-1R resistance (TIR) domain–containing adapter-inducing IFN-β (TRIF) is a Toll-like receptor (TLR) adapter that mediates MyD88-independent induction of type I interferons through activation of IFN regulatory factor 3 and NFκB. We have examined peptides derived from the TRIF TIR domain for ability to inhibit TLR4. In addition to a previously identified BB loop peptide (TF4), a peptide derived from putative helix B of TRIF TIR (TF5) strongly inhibits LPS-induced cytokine and MAPK activation in wild-type cells. TF5 failed to inhibit LPS-induced cytokine and kinase activation in TRIF-deficient immortalized bone-marrow–derived macrophage, but was fully inhibitory in MyD88 knockout cells. TF5 does not block macrophage activation induced by TLR2, TLR3, TLR9, or retinoic acid-inducible gene 1/melanoma differentiation-associated protein 5 agonists. Immunoprecipitation assays demonstrated that TF4 binds to TLR4 but not TRIF-related adaptor molecule (TRAM), whereas TF5 binds to TRAM strongly and TLR4 to a lesser extent. Although TF5 prevented coimmunoprecipitation of TRIF with both TRAM and TLR4, site-directed mutagenesis of the TRIF B helix residues affected TRIF–TRAM coimmunoprecipitation selectively, as these mutations did not block TRIF–TLR4 association. These results suggest that the folded TRIF TIR domain associates with TRAM through the TRIF B helix region, but uses a different region for TRIF–TLR4 association. The B helix peptide TF5, however, can associate with either TRAM or TLR4. In a mouse model of TLR4-driven inflammation, TF5 decreased plasma cytokine levels and protected mice from a lethal LPS challenge. Our data identify TRIF sites that are important for interaction with TLR4 and TRAM, and demonstrate that TF5 is a potent TLR4 inhibitor with significant potential as a candidate therapeutic for human sepsis.Toll-like receptors (TLRs) initiate innate immune responses by recognizing specific pathogen-associated molecules; for example, TLR4 recognizes lipopolysaccharides (LPSs) of Gram-negative bacteria (1, 2). Ligand recognition induces dimerization of cytoplasmic Toll/IL-1R resistance (TIR) domains of two receptor molecules and causes recruitment of intracellular TIR domain-containing adapters. Four adapter proteins participate in TLR4 signaling: myeloid differentiation factor 88 (MyD88) (3), TIR domain-containing adapter protein, also known as MyD88-adapter-like (TIRAP–Mal) (4, 5), TIR domain–containing adapter-inducing IFN-β, also known as TLR adaptor molecule 1 (TRIF–TICAM-1) (6, 7), and TRIF-related adaptor molecule also known as TLR adaptor molecule 2 (TRAM–TICAM-2) (8, 9). TIRAP–Mal is important for MyD88 recruitment to the signaling complex located at the plasma membrane to initiate early NF-κB and mitogen-activated protein kinase (MAPK) activation and induce “MyD88-dependent” proinflammatory cytokines, such as TNF-α and IL-1β (4, 5, 10). TRAM is important for TRIF recruitment to the endosomally located TLR4 signaling complexes to activate IFN regulatory factor 3 (IRF3) and induce IRF3-dependent cytokines, such as IFN-β and RANTES (regulated upon activation normal T-cell expressed and secreted) (8, 9, 11).A typical TIR domain consists of the central five stranded parallel β sheets (the strands are designated as βA–βE) surrounded by 5 α-helices (i.e., αA–αE) (12, 13). The TIRAP–Mal TIR domain has an atypical fold compared with other resolved mammalian TIR structures in that the position of its β-strand B is shifted by 12–18 amino acids toward the C terminus, so that TIRAP TIR does not have a helix B but has an unusually long AB loop (14, 15). Structures of the TIR domains of TLR4, TRIF, and TRAM have not been yet resolved. The TIR domain is a key structural feature present in all TLRs and TLR adapter proteins. TIR domains mediate transient homotypic or heterotypic protein interactions required for agonist-driven assembly of TLR signaling complexes (13, 16, 17). Multiple interactions of TIR domains of TLRs and TLR adapters are required to mediate adapter recruitment and stabilize initial complex (1820). It has been proposed that TLR4 activation leads to formation of several compositionally distinct complexes. Kagan et al. proposed that TLR4 engages TIRAP–MyD88 and TRAM–TRIF sequentially at distinct cellular locations (11), thus implying that the two sets of adapters may compete for the same binding site at the TLR4 homodimer. However, it remains unclear how exactly the four adapters interact with each other and TLR4 to orchestrate TLR4 signaling.The presumed mechanism of signaling inhibition by a decoy peptide is that the peptide competes with its prototype protein for the prototype’s docking site and thereby prevents a protein–protein interaction required for signaling (19). In this study, we have examined cell-permeable decoy peptides derived from the TIR domain of TRIF. Two peptides, TF4 and TF5, from the second loop (BB loop) and the second helical region (helix B) of the TRIF TIR, respectively, potently inhibited LPS-induced activation of MAPKs and induction of MyD88-dependent and TRIF-dependent cytokines in wild-type macrophages. TF5 did not inhibit TLR4 signaling in TRIF−/− immortalized bone-marrow–derived macrophages (iBMDMs) but did exhibit full activity in the MyD88−/− cells. TF5 inhibits TLR4-driven macrophage signaling at a lower dose in vitro compared with TF4 and binds to both TRAM and TLR4, whereas TF4 targets TLR4 but not the TRAM TIR. In a mouse model of TLR4-driven inflammation, TF5 potently decreased the systemic cytokine levels induced in mice by a sublethal LPS dose, and dramatically improved survival of mice challenged with a lethal LPS dose.  相似文献   

2.
3.
Homotypic and heterotypic interactions between Toll/interleukin-1 receptor (TIR) domains in Toll-like receptors (TLRs) and downstream adaptors are essential to evoke innate immune responses. However, such oligomerization properties present intrinsic difficulties in structural studies of TIR domains. Here, using BB-loop mutations that disrupt homotypic interactions, we determined the structures of the monomeric TIR domain-containing adaptor molecule (TICAM)-1 and TICAM-2 TIR domains. Docking of the monomeric structures, together with yeast two hybrid-based mutagenesis assays, reveals that the homotypic interaction between TICAM-2 TIR is indispensable to present a scaffold for recruiting the monomeric moiety of the TICAM-1 TIR dimer. This result proposes a unique idea that oligomerization of upstream TIR domains is crucial for binding of downstream TIR domains. Furthermore, the bivalent nature of each TIR domain dimer can generate a large signaling complex under the activated TLRs, which would recruit downstream signaling molecules efficiently. This model is consistent with previous reports that BB-loop mutants completely abrogate downstream signaling.The extracellular domain of toll-like receptor 4 (TLR4) specifically binds lipopolysaccharides (LPSs) from Gram-negative bacteria, inducing dimerization and leading to the dimerization of cytosolic Toll/interleukin-1 receptor (TIR) domains. This activated conformation of TLR4 recruits the TIR domain of a downstream adaptor molecule, TIR domain-containing adaptor molecule-2 (TICAM-2) [also known as TRIF-related adaptor molecule (TRAM)], that subsequently recruits the TIR domain of another adaptor molecule, TIR domain-containing adaptor molecule-1 (TICAM-1) [also known as TIR domain-containing adaptor inducing IFN-β (TRIF)] (13) at endosomes. Eventually this process activates IFN response factors and generates type-I interferons (IFNs) (47). Elucidation of the homotypic and heterotypic interactions between TICAM-1 and TICAM-2 is essential for understanding of TLR4-mediated type-I IFN generation (8).A large number of TIR domain structures, including receptors and adaptors, have been determined by X-ray crystallography and NMR. The receptors include TLR1 (9), TLR2 (10), and IL-1R accessory protein-like (IL-1RAPL) (11). Adaptors include myeloid differentiation factor 88 (MyD88) (12) and MyD88 adaptor-like (Mal) (13, 14). In addition, AtTIR (15, 16) derived from Arabidopsis thaliana and PdTIR (17) from bacteria have been solved. Each of these TIR domain structures has a ferredoxin fold with five β-strands (βA–βE), five α-helices (αA–αE), and loops connecting β-strands and α-helices (9). Although homotypic interactions of the TIR domains have been proposed based on the crystal structures, most proposed models have small interacting surfaces, possibly due to crystal contacts. Recently, however, a crystal structure of the TLR10 TIR domain was reported that forms a homotypic dimer mediated by the loop connecting βB and αB (designated “BB-loop”) (18). Interestingly, BB-loop mutations in TLR4 were reported to be dominant-negative and abrogated downstream signaling (19). TICAM-1 and TICAM-2 harboring BB-loop mutations are also dominant-negative and unable to form homotypic interactions (1, 2), reinforcing the importance of BB-loop–mediated homotypic dimer formation in signal propagation.Despite extensive structural studies, it is not known why homotypic interactions are essential for downstream signaling (2027). To address this issue, it is necessary to discriminate residues required for homotypic and those required for heterotypic interactions. Here, we first determine the structures of the monomeric BB-loop mutants of the TICAM-1 and TICAM-2 TIR domains using NMR. Then, based on the solution structures of the BB-loop mutants, coupled mutagenesis/yeast two-hybrid experiments, and restrained docking calculations, we show that the homotypic interaction of TICAM-2 TIR is essential to form a scaffold for recruiting the TICAM-1 TIR domain.  相似文献   

4.
5.
Lipopolysaccharide (LPS), also known as endotoxin, activates the innate immune response through toll-like receptor 4 (TLR4) and its coreceptor, MD-2. MD-2 has a unique hydrophobic cavity that directly binds to lipid A, the active center of LPS. Tetraacylated lipid IVa, a synthetic lipid A precursor, acts as a weak agonist to mouse TLR4/MD-2, but as an antagonist to human TLR4/MD-2. However, it remains unclear as to how LPS and lipid IVa show agonistic or antagonistic activities in a species-specific manner. The present study reports the crystal structures of mouse TLR4/MD-2/LPS and TLR4/MD-2/lipid IVa complexes at 2.5 and 2.7 Å resolutions, respectively. Mouse TLR4/MD-2/LPS exhibited an agonistic “m”-shaped 2:2:2 complex similar to the human TLR4/MD-2/LPS complex. Mouse TLR4/MD-2/lipid IVa complex also showed an agonistic structural feature, exhibiting architecture similar to the 2:2:2 complex. Remarkably, lipid IVa in the mouse TLR4/MD-2 complex occupied nearly the same space as LPS, although lipid IVa lacked the two acyl chains. Human MD-2 binds lipid IVa in an antagonistic manner completely differently from the way mouse MD-2 does. Together, the results provide structural evidence of the agonistic property of lipid IVa on mouse TLR4/MD-2 and deepen understanding of the ligand binding and dimerization mechanism by the structurally diverse LPS variants.Toll-like receptors (TLRs) recognize and respond to diverse pathogenic components of microorganisms and provide the first line of defense against microbial infection (1, 2). Among the microbial components, endotoxic lipopolysaccharide (LPS) from a membrane component of Gram-negative bacteria elicits the potent innate immune response through the receptor complex of TLR4 and MD-2 (3, 4). Excessive exposure to LPS often causes exaggerated signaling via TLR4 and fatal septic shock (5, 6), which is associated with a high mortality (20–30%) and is the most common cause of death in intensive care units (5, 6).The lipid A moiety of LPS, which anchors LPS to the outer membrane of Gram-negative bacteria, is responsible for the immunostimulatory activity of LPS (7, 8). Lipid A consists of a 1,4′-bis-phosphorylated diglucosamine backbone to which variable lengths and numbers of acyl chains are covalently linked (8). The two phosphate groups are also important for the agonistic activity of lipid A because deletion of either phosphate group reduces the endotoxic activity (9, 10).TLR4 is a type I transmembrane protein composed of 22 extracellular leucine-rich repeats (LRRs), a transmembrane domain, and the Toll/IL-1 receptor domain (TIR domain) that is essential for TLR signaling and conserved among members of the Toll receptor family (1). TLR4 alone does not directly bind LPS and requires the coreceptor MD-2 (11). MD-2 is associated with the extracellular domain of TLR4 and is indispensable for LPS recognition (4). A member of the MD-2–related lipid-recognition protein family (12), MD-2 directly binds to LPS in its hydrophobic cavity with high affinity (13).Recently, the crystal structure of human TLR4/MD-2/Ra-LPS (Ra chemotype of Escherichia coli LPS) complex (14) was solved, which revealed that five of the six acyl chains of LPS are buried inside the MD-2 cavity. The sixth acyl chain lies on the surface of MD-2, partially exposed to the solvent. Together with the hydrophobic residues of MD-2, the partially exposed acyl chain constitutes the secondary binding site for the hydrophobic patch on the C-terminal convex face of the horseshoe structure of TLR4, leading to the formation of the “m”-shaped 2:2:2 hTLR4/MD-2/LPS complex. The close proximity of the C terminus of the extracellular domain in the complex induced by binding to LPS may allow for dimerization and signaling by the intracellular TIR domains (15, 16).The number and length of the acyl chains determine the agonistic property of lipid A (1719). E. coli lipid A is usually hexaacylated and acts as a potent agonist for all mammalian cells. In contrast, tetraacylated lipid IVa, the precursor of E. coli LPS, acts as an agonist only for some mammalian species. In particular, it acts as a weak agonist on mouse and as an antagonist on human cells (20, 21). Although several studies have investigated the species-specific activity of lipid IVa (2228), these studies primarily used mutational and computational simulation methods. Structural information on the agonistic form of TLR4/MD-2 is limited to the hTLR4/MD-2/LPS complex; no structures of mTLR4/MD-2 complexed with LPS or lipid IVa are currently available. Structural knowledge may provide critical clues regarding the agonistic and antagonistic mechanisms by LPS and lipid IVa ligands that underlie species specificity.Here, we present the two agonistic structures of mouse TLR4/MD-2/Re-LPS (Re chemotype of E. coli LPS) and TLR4/MD-2/lipid IVa complexes at 2.5 and 2.7 Å resolutions, respectively. This structural study will provide better understanding of the LPS recognition and signaling mechanism and will contribute to the development of therapeutic antiseptic shock drugs targeting TLR4/MD-2.  相似文献   

6.
Toll-like receptor 8 (TLR8) senses single-stranded RNA (ssRNA) and initiates innate immune responses. TLR8 requires proteolytic cleavage at the loop region (Z-loop) between leucine-rich repeat (LRR) 14 and LRR15 for its activation. However, the molecular basis of Z-loop processing remains unknown. To elucidate the mechanism of Z-loop processing, we performed biochemical and structural studies of how the Z-loop affects the function of TLR8. TLR8 with the uncleaved Z-loop is unable to form a dimer, which is essential for activation, irrespective of the presence of agonistic ligands. Crystallographic analysis revealed that the uncleaved Z-loop located on the ascending lateral face prevents the approach of the dimerization partner by steric hindrance. This autoinhibition mechanism of dimerization by the Z-loop might be occurring in the proteins of the same subfamily, TLR7 and TLR9.Toll-like receptors (TLRs) constitute a family of innate immune receptors that recognize pathogen-associated molecular patterns (1). The TLR molecule is a type I transmembrane protein characterized by an extracellular leucine-rich repeat (LRR) domain, a transmembrane helix, and an intracellular Toll/interleukin-1 receptor (TIR) homology domain (2). The typical TLR molecule is considered to be monomeric in the absence of ligands, transforming into an activated dimer form on ligand binding, which allows for dimerization of the intracellular TIR domain and subsequent signaling (2).The TLR subfamily comprising TLR7, TLR8, and TLR9 recognizes single-stranded (ss) nucleic acids from viruses and bacteria (3). Specifically, TLR7 and TLR8 recognize uridine- and guanosine-rich single-stranded RNA (ssRNA) (411), whereas TLR9 recognizes ssDNA containing the unmethylated cytosine-phosphate-guanine (CpG) dideoxynucleotide motif (12). Furthermore, TLR7 and TLR8 are also activated by synthetic chemical compounds (13, 14), such as imiquimod (TLR7-specific), resiquimod (R848; both TLR7 and TLR8), and CL075 (both TLR7 and TLR8).Certain regulation mechanisms of the functions of the TLR7–9 subfamily members are shared because of a high degree of sequence similarities (3). They reside on the endosomal membrane, and their transportation from endoplasmic reticulum (ER) to endolysosomes is mediated by the ER membrane protein Unc93B1 (15). Moreover, TLR7–9 possess a long inserted loop region (Z-loop), consisting of ∼30 amino acid residues, between LRR14 and LRR15, and the processing by proteolytic cleavage at the Z-loop is believed to be indispensable for their function (1621). Specifically, the processing at the Z-loop of human TLR8 mediated by furin-like proprotein convertase and cathepsins produces functional TLR8 capable of ligand binding and signaling in endolysosomes. In addition, the cleaved form of TLR8 has been found to be predominant in immune cells (16). Recent structural studies demonstrate that the N- and C-terminal halves of TLR8 after Z-loop cleavage associate with each other, and that both fragments are cooperatively involved in ligand binding (22). Moreover, a recent study revealed that the latter half of the cleaved Z-loop interacts with LRRs to stabilize the TLR8 structure and contributes to ssRNA recognition by TLR8 (23).Although accumulating evidence illustrates the functional importance of Z-loop processing at the cellular level, mechanistic insights into this processing in the regulation of TLR8 function at the molecular level are lacking. Here, to unveil the mechanistic role of Z-loop processing of TLR8, we present the results of a combined structural and biochemical investigation of TLR8 with the uncleaved Z-loop.  相似文献   

7.
Toll-like receptors (TLRs) are crucial in innate recognition of invading micro-organisms and their subsequent clearance. Bacteria are not passive bystanders and have evolved complex evasion mechanisms. Staphylococcus aureus secretes a potent TLR2 antagonist, staphylococcal superantigen-like protein 3 (SSL3), which prevents receptor stimulation by pathogen-associated lipopeptides. Here, we present crystal structures of SSL3 and its complex with TLR2. The structure reveals that formation of the specific inhibitory complex is predominantly mediated by hydrophobic contacts between SSL3 and TLR2 and does not involve interaction of TLR2–glycans with the conserved LewisX binding site of SSL3. In the complex, SSL3 partially covers the entrance to the lipopeptide binding pocket in TLR2, reducing its size by ∼50%. We show that this is sufficient to inhibit binding of agonist Pam2CSK4 effectively, yet allows SSL3 to bind to an already formed TLR2–Pam2CSK4 complex. The binding site of SSL3 overlaps those of TLR2 dimerization partners TLR1 and TLR6 extensively. Combined, our data reveal a robust dual mechanism in which SSL3 interferes with TLR2 activation at two stages: by binding to TLR2, it blocks ligand binding and thus inhibits activation. Second, by interacting with an already formed TLR2–lipopeptide complex, it prevents TLR heterodimerization and downstream signaling.In recent years, Staphylococcus aureus has become a major health threat to both humans and domestic animals. It is found as a commensal bacterium in ∼30% of the human population, but when it becomes infectious it can cause a wide diversity of diseases, ranging from mild skin infections to life-threatening invasive conditions such as pneumonia and sepsis (1). Increased antibiotic resistance and a high amount of virulence factors secreted by S. aureus contribute to its emergence as a pathogen. Among these secreted virulence factors are the staphylococcal superantigen-like proteins (SSLs), a family of 14 proteins located on two genomic clusters (24). Recently, we and others identified SSL3 as a potent inhibitor of Toll-like receptor 2 (TLR2) (5, 6), an innate immunity receptor that is a dominant factor in immune recognition of S. aureus (710).TLR2 belongs to a family of 10 homologous innate immunity receptors that are activated by pathogen-associated molecular patterns (PAMPs) (11). TLR2 binds bacterial lipopeptides and lipoproteins. Subsequent formation of heterodimers with TLR1 or TLR6 leads to MyD88-dependent activation of the NF-κB pathway (12). TLR2 has dual ligand specificity that is determined by its dimerization partner; stimulation by diacyl lipopeptides from Gram-positive bacteria, including S. aureus, induces the formation of heterodimers with TLR6 (13), whereas triacyl lipopeptides from Gram-negative bacteria initiate formation of TLR2–TLR1 dimers (14). The structural basis for lipopeptide specificity was revealed by crystal structures of TLR2–TLR1 and TLR2–TLR6 complexes with their respective lipopeptide analogs Pam3CSK4 and Pam2CSK4: TLR2 binds two lipid tails in a large hydrophobic pocket, whereas the third lipid tail of triacyl lipopeptides is accommodated by a smaller pocket present in TLR1, but not in TLR6 (15, 16).The family of SSL proteins, including SSL3, share structural similarities to superantigens, but lack superantigenic activity. Interestingly, the functions that have been discovered for SSLs so far have all been linked to immune evasion. SSL5 inhibits neutrophil extravasation (17, 18) and phagocyte function (19, 20), SSL7 binds IgA and inhibits complement (21), and SSL10 inhibits IgG1-mediated phagocytosis (22, 23), blood coagulation (24), and the chemokine receptor CXCR4 (25). In addition to SSL3, also weak TLR2 inhibitory activity was observed for SSL4 (5), but it remains unknown whether that is its dominant function. This variety of immunomodulatory molecules and functions reflects the importance of the different components of our innate immune system in the defense against S. aureus (26).In this study we determined the crystal structures of SSL3 and the SSL3–TLR2 complex. In combination with mutagenesis and binding studies, our data provide a novel working mechanism of a functional TLR2 antagonist.  相似文献   

8.
Systemic lupus erythematosus (SLE) is a complex autoimmune disease with diverse clinical presentations characterized by the presence of autoantibodies to nuclear components. Toll-like receptor (TLR)7, TLR8, and TLR9 sense microbial or endogenous nucleic acids and are implicated in the development of SLE. In mice TLR7-deficiency ameliorates SLE, but TLR8- or TLR9-deficiency exacerbates the disease because of increased TLR7 response. Thus, both TLR8 and TLR9 control TLR7 function, but whether TLR8 and TLR9 act in parallel or in series in the same or different cell types in controlling TLR7-mediated lupus remains unknown. Here, we reveal that double TLR8/9-deficient (TLR8/9−/−) mice on the C57BL/6 background showed increased abnormalities characteristic of SLE, including splenomegaly, autoantibody production, frequencies of marginal zone and B1 B cells, and renal pathology compared with single TLR8−/− or TLR9−/− mice. On the cellular level, TLR8−/− and TLR8/9−/− dendritic cells were hyperesponsive to TLR7 ligand R848, but TLR9−/− cells responded normally. Moreover, B cells from TLR9−/− and TLR8/9−/− mice were hyperesponsive to R848, but TLR8−/− B cells were not. These results reveal that TLR8 and TLR9 have an additive effect on controlling TLR7 function and TLR7-mediated lupus; however, they act on different cell types. TLR8 controls TLR7 function on dendritic cells, and TLR9 restrains TLR7 response on B cells.Systemic lupus erythematosus (SLE) is a complex chronic autoimmune disease that arises spontaneously and is characterized by production of autoantibodies against self-nucleic acids and associated proteins (1). These autoantibodies bind self-nucleic acids released by dying cells and form immune complexes that accumulate in different parts of the body, leading to inflammation and tissue damage. The kidneys, skin, joints, lungs, serous membranes, as well as, the cardiovascular, nervous and musculoskeletal system become targets of inflammation at onset or during the course of the disease (2). The etiology of SLE is unknown, yet genetics, sex, infectious agents, environmental factors, and certain medications may play a role in the initiation of the disease by causing alterations in lymphoid signaling, antigen presentation, apoptosis, and clearance of immune complexes (3, 4).Toll-like receptors (TLRs) detect specific microbial components widely expressed by bacteria, fungi, protozoa, and viruses, and initiate signaling pathways critical for induction of immune responses to infection (5). In contrast to the cell surface TLRs that detect bacterial cell wall components and viral particles, nucleic acid-sensing TLRs are localized mainly within endosomal compartments (6). Human endosomal TLRs consist of TLR3, which senses viral double-stranded RNA (dsRNA) (7), TLR7 and TLR8, which recognize viral single-stranded RNA (810), and TLR9, which detects bacterial and viral unmethylated CpG-containing DNA motifs (11). Interestingly, these endosomal TLRs are also able to detect self-nucleic acids (1214). Although the endosomal localization isolate TLR3, TLR7, TLR8, and TLR9 away from self-nucleic acids in the extracellular space, still self-RNA or -DNA can become a potent trigger of cell activation when transported into TLR-containing endosomes, and such recognition can result in sterile inflammation and autoimmunity, including SLE (4, 15, 16). The connection of the endosomal TLRs with SLE originates mainly from mouse models, where TLR7 signaling seems to play a central role. TLR7 gene duplication is the cause for the development of lupus in mice bearing the Y chromosome-linked autoimmune accelerating (Yaa) locus that harbors 17 genes, including TLR7 (17, 18). In TLR7 transgenic mouse lines, a modest increase in TLR7 expression promotes autoreactive lymphocytes with RNA specificities and myeloid cell proliferation, but a substantial increase in TLR7 expression causes fatal acute inflammatory pathology and profound dendritic cell (DC) dysregulation (17). In addition, studies in several lupus-prone mouse strains have revealed that TLR7-deficiency ameliorates disease, but TLR9-deficiency exacerbates it. Interestingly, this controversy can be explained by the enhanced TLR7 activity in the TLR9-deficient lupus mice (19, 20). Although murine TLR8 does not seem so far to be able to sense a ligand (21, 22), we have shown previously that it plays an important biological role in controlling TLR7-mediated lupus. Indeed, TLR8-deficiency in mice (on the C57BL/6 background that is not prone to lupus) leads to lupus development because of increased TLR7 expression and signaling in DCs (23). Thus, tight control and regulation of TLR7 is pivotal for avoiding SLE and inflammatory pathology in mice. Recent studies in humans have also revealed that increased expression of TLR7 is associated with increased risk for SLE (2426).Nucleic acid TLRs are expressed in many cell types, including DCs, plasmacytoid DCs (pDCs) and B cells, all of which play a central role in SLE development. TLR7, TLR8, and TLR9 signal through the adaptor molecule myeloid differentiation primary response gene 88 (MyD88), whereas TLR3 signals via the adaptor TRIF (Toll/IL-1 receptor domain-containing adaptor inducing IFN-β) (5). MyD88-deficiency abrogates most attributes of lupus in several lupus-prone mouse strains (19, 2729). Moreover, deficiency for Unc93B1, a multipass transmembrane protein that controls trafficking of TLRs from the endoplasmic reticulum to endolysosomes and is required for nucleic acid-sensing TLR function (30), also abrogates many clinical parameters of disease in mouse lupus strains, suggesting that endosomal TLRs are critical in this disease (31). Interestingly, TLR9 competes with TLR7 for Unc93B1-dependent trafficking and predominates over TLR7 (32). TLR9 predominance is reversed to TLR7 by a D34A mutation in Unc93B1 and mice that carry this mutation show TLR7-dependent, systemic lethal inflammation (32).Thus, in mice both TLR8 and TLR9 control TLR7-mediated lupus, but it is unknown if these TLRs act in parallel or in series in the same or different cell types and if they have an additive effect or not in controlling TLR7. To address these issues, we generated double TLR8/TLR9-deficient (TLR8/9−/−) mice and analyzed and compared the lupus phenotype in TLR8−/−, TLR9−/−, and TLR8/9−/− mice. Our data revealed that TLR8/9−/− mice have increased abnormalities characteristic of SLE and that both TLR8 and TLR9 keep TLR7-mediated lupus under control, but they act in different cell types. On DCs TLR7 function is ruled by TLR8, whereas on B cells TLR7 is mastered by TLR9.  相似文献   

9.
10.
11.
12.
Systemic sclerosis (SSc) is a multisystem autoimmune disorder with clinical manifestations resulting from tissue fibrosis and extensive vasculopathy. A potential disease susceptibility gene for SSc is IFN regulatory factor 5 (IRF5), whose SNP is associated with milder clinical manifestations; however, the underlying mechanisms of this association remain elusive. In this study we examined IRF5-deficient (Irf5−/−) mice in the bleomycin-treated SSc murine model. We show that dermal and pulmonary fibrosis induced by bleomycin is attenuated in Irf5−/− mice. Interestingly, we find that multiple SSc-associated events, such as fibroblast activation, inflammatory cell infiltration, endothelial-to-mesenchymal transition, vascular destabilization, Th2/Th17 skewed immune polarization, and B-cell activation, are suppressed in these mice. We further provide evidence that IRF5, activated by Toll-like receptor 4 (TLR4), binds to the promoters of various key genes involved in SSc disease pathology. These observations are congruent with the high level of expression of IRF5, TLR4, and potential endogenous TLR4 ligands in SSc skin lesions. Our study sheds light on the TLR4-IRF5 pathway in the pathology of SSc with clinical implications of targeting the IRF5 pathways in the suppression of disease development.Systemic sclerosis (SSc) is a multisystem connective tissue disease characterized by immune abnormalities, vasculopathy, and extensive tissue fibrosis (1). Based on the results of etiological and genetic studies, the conventional wisdom is that SSc is caused by a complex interplay between genetic factors and environmental influences. For instance, the biggest risk factor for SSc is family history (2). On the other hand, concordance for SSc is around 5% in twins and is similar in monozygotic and dizygotic twins, whereas antinuclear antibodies are detected more frequently in the healthy monozygotic twin sibling than in the healthy dizygotic twin sibling of an SSc patient (3). In addition, most SSc susceptibility genes are HLA haplotypes and non-HLA immune-related genes that are shared by other collagen diseases (4). Therefore, genetic factors are likely associated with autoimmunity, increasing the susceptibility to autoimmune diseases including SSc, but additional environmental factors are required to induce clinically definite SSc in genetically predisposed individuals. Despite these etiological and genetic data, the entire process of the SSc development and pathogenesis remains elusive.Therefore it is important to elucidate the molecular mechanism(s) underlying SSc pathogenesis. In this regard, much attention has been focused recently on the innate immune signaling via Toll-like receptors (TLRs) in various pathological conditions. For instance, fibroblasts and endothelial cells in SSc lesional skin highly express TLR4, originally identified as the receptor for bacterial LPS, and TLR4 signaling amplifies the sensitivity to TGF-β in dermal fibroblasts (57). It also was shown that dermal and lung fibrosis is attenuated in bleomycin (BLM)-treated TLR4-deficient mice (7). Endogenous potential TLR4 ligands are up-regulated in SSc lesional skin (57), and serum levels correlate with severe organ involvement and immunological abnormalities (8, 9). Therefore, the TLR4 signaling pathway is suspected to play a central role in the SSc pathogenesis.Although how the TLR4 signaling pathway contributes to SSc pathogenesis remains enigmatic, it is interesting that several independent case-control and genome-wide association studies identify IFN regulatory factor 5 (IRF5), a member of the IFN regulatory factor (IRF) family, as an SSc susceptibility gene (1015). IRFs were identified primarily in the research of the type I IFN system and have been shown to have functionally diverse roles in the regulation of the innate and adaptive immune responses (16). Reflecting such property of IRFs, SNPs of IRFs have been linked to the development of various immune and inflammatory disorders. IRF5 is of particular interest, being implicated in multiple sclerosis, systemic lupus erythematosus, rheumatoid arthritis, and SSc (14). Thus far an association of certain SNPs within the IRF5 promoter with the risk and severity of SSc has been reported (1015), but whether and how IRF5 is activated to contribute to disease development remains unknown.Stimulation of TLRs triggers the activation of myeloid differentiation factor 88 (MyD88)-dependent and/or independent pathways (16). IRF5 is activated via the MyD88 pathway in dendritic cells and macrophages (17). TLR-activated IRF5 mediates the induction of genes IL-6, IL-12, and TNF-α (17). Hence, an intriguing possibility is that TLR4-mediated activation of IRF5 is involved in SSc. We therefore studied the role of IRF5 in the regulation of genes associated with the susceptibility to and the severity of SSc using IRF5-deficient mice in the context of TLR4 signaling. We show that IRF5, activated by TLR4, binds to the promoters of various key genes involved in the disease symptoms. We discuss our findings in terms of the complexity of SSc and its clinical implications.  相似文献   

13.
Dendritic cells (DCs) phagocytose large particles like bacteria at sites of infection and progressively degrade them within maturing phagosomes. Phagosomes in DCs are also signaling platforms for pattern recognition receptors, such as Toll-like receptors (TLRs), and sites for assembly of cargo-derived peptides with major histocompatibility complex class II (MHC-II) molecules. Although TLR signaling from phagosomes stimulates presentation of phagocytosed antigens, the mechanisms underlying this enhancement and the cell surface delivery of MHC-II–peptide complexes from phagosomes are not known. We show that in DCs, maturing phagosomes extend numerous long tubules several hours after phagocytosis. Tubule formation requires an intact microtubule and actin cytoskeleton and MyD88-dependent phagosomal TLR signaling, but not phagolysosome formation or extensive proteolysis. In contrast to the tubules that emerge from endolysosomes after uptake of soluble ligands and TLR stimulation, the late-onset phagosomal tubules are not essential for delivery of phagosome-derived MHC-II–peptide complexes to the plasma membrane. Rather, tubulation promotes MHC-II presentation by enabling maximal cargo transfer among phagosomes that bear a TLR signature. Our data show that phagosomal tubules in DCs are functionally distinct from those that emerge from lysosomes and are unique adaptations of the phagocytic machinery that facilitate cargo exchange and antigen presentation among TLR-signaling phagosomes.Professional phagocytes take up large particles, such as bacteria, by phagocytosis and submit them to an increasingly harsh environment during phagosome maturation (1). Phagocytes concomitantly alert the immune system that an invader is present via signaling programs initiated by pattern recognition receptors, such as Toll-like receptors (TLRs) (2). Conventional dendritic cells (DCs) also alter and optimize phagosome maturation and TLR-signaling programs to preserve bacterial antigens for loading onto MHC class I and class II (MHC-II) molecules and optimize cytokine secretion to stimulate and direct T-cell responses to the invading agent (3, 4). DC presentation of soluble antigen is facilitated by TLR-driven tubulation of lysosomes that harbor MHC-II–peptide complexes and by consequent fusion of tubulovesicular structures with the plasma membrane (57); however, little is known about the mechanism by which signaling pathways influence the formation or presentation of phagosome-derived MHC-II–peptide complexes, key processes in the adaptive immunity to bacterial pathogens.TLRs respond to microbial ligands at the plasma membrane and in intracellular stores (8). TLR stimulation at the plasma membrane, endosomes, or phagosomes elicits distinct signaling pathways via two sets of adaptors, TIRAP (or MAL)-MyD88 and TRAM-TRIF (8, 9), which induce proinflammatory cytokine secretion and other downstream responses. TLRs such as TLR2 and TLR4 are recruited to macrophage and DC phagosomes at least partly from an intracellular pool (1013), and signal autonomously from phagosomes independent of plasma membrane TLRs (11, 14, 15). Autonomous phagosomal signaling from TLRs or Fcγ receptors enhances the degradation of phagocytosed proteins and assembly of MHC-II with their derived peptides (1416). Phagosomal TLR signaling has been proposed to also promote the reorganization of phagosome-derived MHC-II-enriched compartments (MIICs) to favor the delivery of MHC-II–peptide complexes to the plasma membrane (17), analogous to TLR-stimulated formation of tubules from MIICs/lysosomes (1820) that fuse with the plasma membrane (7) and extend toward the immunologic synapse with T cells (5). Tubules emerge from phagosomes in macrophages shortly after phagocytosis and likely function in membrane recycling during early phagosome maturation stages (2123), but tubules at later stages that might facilitate the presentation of phagosome-derived MHC-II–peptide complexes have not been reported previously. Moreover, a role for TLR signaling in formation of phagosome-derived tubules has not been established.Herein we show that in DCs, maturing phagosomes undergo extensive tubulation up to several hours after phagocytosis, and that tubulation requires TLR and MyD88 signaling and an intact actin and microtubule cytoskeleton. Unlike lysosome tubulation, phagosome tubulation is not essential for MHC-II–peptide transport to the cell surface. Rather, it contributes to content exchange among phagosomes that carry a TLR signature, and thereby enhances presentation of phagocytosed antigens from potential pathogens.  相似文献   

14.
The proper trafficking and localization of Toll-like receptors (TLRs) are important for specific ligand recognition and efficient signal transduction. The TLRs sensing bacterial membrane components are expressed on the cell surface and recruit signaling adaptors to the plasma membrane upon stimulation. On the contrary, the nucleotide-sensing TLRs are mostly found inside cells and signal from the endolysosomes in an acidic pH-dependent manner. Trafficking of the nucleotide-sensing TLRs from the endoplasmic reticulum to the endolysosomes strictly depends on UNC93B1, and their signaling is completely abolished in the 3d mutant mice bearing the H412R mutation of UNC93B1. In contrast, UNC93B1 was considered to have no role for the cell surface-localized TLRs and signaling via TLR1, TLR2, TLR4, and TLR6 is normal in the 3d mice. Unexpectedly, we discovered that TLR5, a cell surface receptor for bacterial protein flagellin, also requires UNC93B1 for plasma membrane localization and signaling. TLR5 physically interacts with UNC93B1, and the cells from the 3d or UNC93B1-deficient mice not only lack TLR5 at the plasma membrane but also fail to secret cytokines and to up-regulate costimulatory molecules upon flagellin stimulation, demonstrating the essential role of UNC93B1 in TLR5 signaling. Our study reveals that the role of UNC93B1 is not limited to the TLRs signaling from the endolysosomes and compels the further probing of the mechanisms underlying the UNC93B1-assisted differential targeting of TLRs.Toll-like receptors (TLRs) sense unique microbial structures or host-derived molecules released from stressed or dying cells to initiate the innate immune responses (1). TLRs are composed of three domains: the leucine-rich repeat (LRR) domain responsible for ligand binding, a single transmembrane domain, and the cytoplasmic Toll/IL-1 receptor homology domain by which TLRs recruit adaptor molecules for downstream signal transduction. Activated TLRs stimulate the NF-κB, MAPK, and IFN regulatory factor pathways, leading to the expression of diverse inflammatory cytokines, chemokines, and type I interferons. TLRs also activate antigen presenting cells to induce costimulatory molecules and coordinate various aspects of adaptive immune responses (2).The members of the TLR family can be classified into two groups based on their subcellular localization patterns (35). TLR1, TLR2, TLR4, and TLR6, which mainly recognize the components of bacterial cell membrane, are located on the cell surface and initiate signaling thereat. In contrast, the nucleotide-sensing TLRs such as TLR3, TLR7, TLR8, TLR9, and TLR13 are largely found in endolysosomes and require an acidic environment for their efficient signaling. Additionally, TLR11 and TLR12, the sensors for Toxoplasma protein profilin, are also expressed inside cells and transmit signals in an acidic pH-dependent manner (68). All the intracellular TLRs commonly bind to a multispanning membrane protein UNC93B1, which is required for their proper localization and signaling (613). One missense mutation (H412R) of UNC93B1, found in a chemically mutagenized mouse strain called 3d, hinders binding of UNC93B1 with TLRs and prevents their exit from the endoplasmic reticulum (ER) (911). Consequently, signaling by all endosomal TLRs is abolished in the cells from 3d mice. In contrast, trafficking and signaling of the cell surface-localized TLRs such as TLR2 and TLR4 are not affected by the UNC93B1 mutation (9, 11).The proper localization of TLRs is critical not only for efficient signaling but also for preventing undesirable receptor hyperactivation (14, 15). Especially, sequestration of the nucleotide-sensing TLRs in endolysosomes significantly contributes to attenuating the immune stimulation by host-derived nucleotides abundant in the extracellular spaces (14). Structural discrimination of microbial vs. mammalian nucleotides is not straightforward, and a mutant TLR9 protein, engineered to artificially localize at the plasma membrane, responds to mammalian DNA as well as the CpG oligonucleotides mimicking bacterial DNA. As a result, mice expressing such mutant TLR9 succumb to systemic autoinflammation and die prematurely (15). Therefore, regulatory mechanisms for localization and trafficking of TLRs need to be tightly controlled.TLR5 recognizes flagellin, the major protein subunit of bacterial flagellum, and functions as a critical innate sensor for flagellated bacteria in all mucous organs (1618). TLR5 plays an important role in intestinal homeostasis mediating the immune adaptation to symbiotic microflora as well as defense against pathogenic bacterial infection (1921). In addition, systemic injection of flagellin confers protection against ionizing radiation in a TLR5-dependent manner, implying that TLR5 agonism might be clinically used for radioprotection (22). TLR5 overexpressed in the intestinal epithelial cells was exclusively found on the basolateral surface, accounting for the selective induction of proinflammatory cytokine by basolateral but not by apical flagellin (17). Also, we recently demonstrated that endogenous TLR5 is expressed at the cell surface of mouse neutrophils, monocytes, and dendritic cells (DCs) in a TLR-specific chaperone PRAT4A-dependnet manner (23). However, other regulatory mechanisms for the localization of TLR5 at the plasma membrane are unknown. Here, we show that UNC93B1 binds to TLR5, travels to the plasma membrane with the receptor, and is required for flagellin-induced signaling at the cell surface.  相似文献   

15.
16.
The ASPP2 (also known as 53BP2L) tumor suppressor is a proapoptotic member of a family of p53 binding proteins that functions in part by enhancing p53-dependent apoptosis via its C-terminal p53-binding domain. Mounting evidence also suggests that ASPP2 harbors important nonapoptotic p53-independent functions. Structural studies identify a small G protein Ras-association domain in the ASPP2 N terminus. Because Ras-induced senescence is a barrier to tumor formation in normal cells, we investigated whether ASPP2 could bind Ras and stimulate the protein kinase Raf/MEK/ERK signaling cascade. We now show that ASPP2 binds to Ras–GTP at the plasma membrane and stimulates Ras-induced signaling and pERK1/2 levels via promoting Ras–GTP loading, B-Raf/C-Raf dimerization, and C-Raf phosphorylation. These functions require the ASPP2 N terminus because BBP (also known as 53BP2S), an alternatively spliced ASPP2 isoform lacking the N terminus, was defective in binding Ras–GTP and stimulating Raf/MEK/ERK signaling. Decreased ASPP2 levels attenuated H-RasV12–induced senescence in normal human fibroblasts and neonatal human epidermal keratinocytes. Together, our results reveal a mechanism for ASPP2 tumor suppressor function via direct interaction with Ras–GTP to stimulate Ras-induced senescence in nontransformed human cells.ASPP2, also known as 53BP2L, is a tumor suppressor whose expression is altered in human cancers (1). Importantly, targeting of the ASPP2 allele in two different mouse models reveals that ASPP2 heterozygous mice are prone to spontaneous and γ-irradiation–induced tumors, which rigorously demonstrates the role of ASPP2 as a tumor suppressor (2, 3). ASPP2 binds p53 via the C-terminal ankyrin-repeat and SH3 domain (46), is damage-inducible, and can enhance damage-induced apoptosis in part through a p53-mediated pathway (1, 2, 710). However, it remains unclear what biologic pathways and mechanisms mediate ASPP2 tumor suppressor function (1). Indeed, accumulating evidence demonstrates that ASPP2 also mediates nonapoptotic p53-independent pathways (1, 3, 1115).The induction of cellular senescence forms an important barrier to tumorigenesis in vivo (1621). It is well known that oncogenic Ras signaling induces senescence in normal nontransformed cells to prevent tumor initiation and maintain complex growth arrest pathways (16, 18, 2124). The level of oncogenic Ras activation influences its capacity to activate senescence; high levels of oncogenic H-RasV12 signaling leads to low grade tumors with senescence markers, which progress to invasive cancers upon senescence inactivation (25). Thus, tight control of Ras signaling is critical to ensure the proper biologic outcome in the correct cellular context (2628).The ASPP2 C terminus is important for promoting p53-dependent apoptosis (7). The ASPP2 N terminus may also suppress cell growth (1, 7, 2933). Alternative splicing can generate the ASPP2 N-terminal truncated protein BBP (also known as 53BP2S) that is less potent in suppressing cell growth (7, 34, 35). Although the ASPP2 C terminus mediates nuclear localization, full-length ASPP2 also localizes to the cytoplasm and plasma membrane to mediate extranuclear functions (7, 11, 12, 36). Structural studies of the ASPP2 N terminus reveal a β–Grasp ubiquitin-like fold as well as a potential Ras-binding (RB)/Ras-association (RA) domain (32). Moreover, ASPP2 can promote H-RasV12–induced senescence (13, 15). However, the molecular mechanism(s) of how ASPP2 directly promotes Ras signaling are complex and remain to be completely elucidated.Here, we explore the molecular mechanisms of how Ras-signaling is enhanced by ASPP2. We demonstrate that ASPP2: (i) binds Ras-GTP and stimulates Ras-induced ERK signaling via its N-terminal domain at the plasma membrane; (ii) enhances Ras-GTP loading and B-Raf/C-Raf dimerization and forms a ASPP2/Raf complex; (iii) stimulates Ras-induced C-Raf phosphorylation and activation; and (iv) potentiates H-RasV12–induced senescence in both primary human fibroblasts and neonatal human epidermal keratinocytes. These data provide mechanistic insight into ASPP2 function(s) and opens important avenues for investigation into its role as a tumor suppressor in human cancer.  相似文献   

17.
G protein-coupled receptor kinase-interactor 2 (GIT2) regulates thymocyte positive selection, neutrophil-direction sensing, and cell motility during immune responses by regulating the activity of the small GTPases ADP ribosylation factors (Arfs) and Ras-related C3 botulinum toxin substrate 1 (Rac1). Here, we show that Git2-deficient mice were more susceptible to dextran sodium sulfate (DSS)-induced colitis, Escherichia coli, or endotoxin-shock challenge, and a dramatic increase in proinflammatory cytokines was observed in Git2 knockout mice and macrophages. GIT2 is a previously unidentified negative regulator of Toll-like receptor (TLR)-induced NF-κB signaling. The ubiquitination of TNF receptor associated factor 6 (TRAF6) is critical for the activation of NF-κB. GIT2 terminates TLR-induced NF-κB and MAPK signaling by recruiting the deubiquitinating enzyme Cylindromatosis to inhibit the ubiquitination of TRAF6. Finally, we show that the susceptibility of Git2-deficient mice to DSS-induced colitis depends on TLR signaling. Thus, we show that GIT2 is an essential terminator of TLR signaling and that loss of GIT2 leads to uncontrolled inflammation and severe organ damage.Inflammatory bowel disease (IBD), which includes the two main forms Crohn’s disease (CD) and ulcerative colitis (UC), is generally thought to develop from an abnormal immune response to the gut luminal microbiota in genetically predisposed individuals (1, 2). The innate immune system, which senses the commensal flora, is critical for intestinal homeostasis and tissue repair after injury (3). Toll-like receptor (TLR) signaling is responsible for the development of colitis (4).Tight control of TLR signaling is crucial for the clearance of pathogens and for avoiding nonresolving inflammation (5, 6). Uncontrolled activation of the TLR-triggered signaling pathway promotes chronic inflammation and autoimmune diseases, such as IBD and obesity (7, 8). TLRs mediate host responses to microorganisms and induce inflammatory responses that are linked to the expression of proinflammatory cytokines and activation of the NF-κB pathway (7). Negative regulation of TLR signaling is crucial to maintain immune homeostasis. The tumor suppressor CYLD is a negative regulator of TLR signaling by deubiquitinating TRAF6 and NEMO (911). A deficiency in the catalytic domain of CYLD in mice causes elevated NF-κB activity, and those mice die shortly after birth (12).GIT2 belongs to the group of ADP ribosylation factor (Arf)-directed GTPase activating proteins (GAPs). More than 24 proteins with Arf GAP domains have been identified in humans (13). To our knowledge, none of these proteins have been reported to directly regulate TLR signaling. GIT2 negatively regulates the activation of Arf GTPases through its N-terminal, Arf GTPase activating protein (GAP) domain (14) and of Rac GTPase through its interaction with PIX (p21-activated kinase-interacting exchange factor) (15). Git2 deficiency leads to spontaneous splenomegaly, hypersusceptibility to infection, increased oxygen anion production by neutrophils, and impaired positive selection of CD4 single-positive thymocytes in thymus (16, 17). GIT2 is highly expressed in monocytes and macrophages, especially after stimulation by LPS (18, 19). Combining its known relationships with novel transeQTLs extends the connections of GIT2 to a host of inflammatory mediators (19). We identified that GIT2 interacted with the components of the NF-κB signaling pathway by a large-scale yeast two-hybrid screening (20). Cumulatively, these results show the important roles of GIT2 in innate and adaptive immunity. However, whether GIT2 directly regulates the intestinal immune response remains unknown, and the genetic evidence is lacking to support its physiological roles during intestinal immune responses by directly regulating the TLR signaling pathway.Here, we show that knockout of Git2 induced severe acute and chronic colitis after dextran sulfate sodium (DSS) treatment. Git2-deficient mice were more susceptible to infection by Escherichia coli and endotoxin shock. Git2 deficiency resulted in a greater TLR-induced production of proinflammatory cytokines in vitro and in vivo. GIT2 inhibited the TLR-induced signaling pathway and was critical in inhibiting the activation of TRAF6. Moreover, GIT2 terminated the TLR signaling pathway by recruiting the deubiquitinating enzyme CYLD to the TRAF6 ubiquitin ligase. Thus, GIT2 is a critical terminator of TLR-induced inflammatory responses.  相似文献   

18.
19.
20.
Dimerization of Toll-like receptor 4 (TLR4)/myeloid differentiation factor 2 (MD2) heterodimers is critical for both MyD88- and TIR-domain–containing adapter-inducing IFN-β (TRIF)-mediated signaling pathways. Recently, Zanoni et al. [(2011) Cell 147(4):868–880] reported that cluster of differentiation 14 (CD14) is required for LPS-/Escherichia coli- induced TLR4 internalization into endosomes and activation of TRIF-mediated signaling in macrophages. We confirmed their findings with LPS but report here that CD14 is not required for receptor endocytosis and downstream signaling mediated by TLR4/MD2 agonistic antibody (UT12) and synthetic small-molecule TLR4 ligands (1Z105) in murine macrophages. CD14 deficiency completely ablated the LPS-induced TBK1/IRF3 signaling axis that mediates production of IFN-β in murine macrophages without affecting MyD88-mediated signaling, including NF-κB, MAPK activation, and TNF-α and IL-6 production. However, neither the MyD88- nor TRIF-signaling pathways and their associated cytokine profiles were altered in the absence of CD14 in UT12- or 1Z105-treated murine macrophages. Eritoran (E5564), a lipid A antagonist that binds the MD2 “pocket,” completely blocked LPS- and 1Z105-driven, but not UT12-induced, TLR4 dimerization and endocytosis. Furthermore, TLR4 endocytosis is induced in macrophages tolerized by exposure to either LPS or UT12 and is independent of CD14. These data indicate that TLR4 receptor endocytosis and the TRIF-signaling pathway are dissociable and that TLR4 internalization in macrophages can be induced by UT12, 1Z105, and during endotoxin tolerance in the absence of CD14.Toll-like receptor 4 (TLR4) signaling plays a crucial role in host defense against Gram-negative bacteria by recognizing the outer membrane component, lipopolysaccharide (LPS) (13). TLR4 signaling is initiated by transfer of an LPS monomer from LPS binding protein (LBP) to cluster of differentiation 14 (CD14) (GPI-linked or soluble). In turn, CD14 transfers monomeric LPS to myeloid differentiation factor 2 (MD-2), a protein that associates noncovalently with TLR4 (4). Appropriate ligand binding to MD2 results in dimerization of two TLR4/MD2 complexes (4). TLR4 is unique in that it is the only TLR that activates both myeloid differentiation primary response 88 (MyD88) and TIR-domain–containing adapter-inducing IFN-β (TRIF)-dependent signaling pathways (5, 6). MyD88-mediated, TLR4 signaling occurs mainly at plasma membranes and involves IL-1R–associated kinases phosphorylation, association of TNF-receptor–associated factor 6, and downstream signaling that results in NF-κB activation and induction of proinflammatory mediators such as TNF-α and IL-6 (7). In contrast, TRIF-mediated signaling in response to LPS occurs at the endosomal membrane after internalization of the TLR4 that, in turn, activates IFN regulatory factor 3 (IRF3), resulting in production of IFN-β, IP-10, and other IRF-3–dependent genes, as well as delayed NF-κB activation (8). Recent studies have shown that the endocytosis of TLR4 is tightly controlled by several molecules. Rab11a, ARF6, and p120-catenin have been implicated in Escherichia coli/LPS-induced TLR4 endocytosis and IRF3 activation (911). Zanoni et al. showed that CD14 plays critical roles in translocation of TLR4 into endosomes and in activation of IRF3 that are dependent upon the enzymatic activities of PLCγ2 and Syk (12). However, CD14-independent TLR4 endocytosis and TRIF signaling have not been reported.UT12 is a monoclonal antibody (MAb) with specificity for the mouse TLR4/MD2 complex and mediates LPS-like signaling (13). It has been shown that UT12 induces endotoxic shock-like symptoms in mice including augmentation of TNF-α and IL-6. Furthermore, UT12 induced long-term tolerance and protection against LPS-induced lethal shock in mice (14). However, the ability of UT12 to induce translocation of TLR4/MD2 into endosomes, as well as its potential for mediating TRIF-dependent signaling, has not been reported. Recently, a group of substituted pyrimido[5,4-b]indoles, synthetic ligands for TLR4 that activate NF-κB that act in a CD14-independent manner, were discovered by high-throughput screening (15). These synthetic ligands induced IL-6 and IP-10 in a TLR4/MD2-dependent, but CD14-independent manner (16). They, too, have not been tested for TLR4 endocytosis and TRIF-dependent intermediates.In this study, we report, for the first time to our knowledge, CD14-independent translocation of TLR4 to endosomes and TRIF signaling by UT12 and small synthetic TLR4 ligands (1Z105). A TLR4 antagonist, Eritoran, that binds to a deep hydrophobic pocket in MD2 and blocks signaling induced by LPS, UT12, and 1Z105, blocked only TLR4 internalization and dimerization induced by LPS and 1Z105. Despite TLR4/MD2 internalization, endotoxin-tolerized macrophages fail to activate TRIF-mediated signaling. These findings reveal previously unidentified insights into the possible role of CD14 in LPS-mediated TLR4 endocytosis and signaling and demonstrate that TLR4 endocytosis and signaling are dissociable processes.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号