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 31 毫秒
1.
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   

2.
3.
Toll-like receptor (TLR) signaling is initiated by dimerization of intracellular Toll/IL-1 receptor resistance (TIR) domains. For all TLRs except TLR3, recruitment of the adapter, myeloid differentiation primary response gene 88 (MyD88), to TLR TIR domains results in downstream signaling culminating in proinflammatory cytokine production. Therefore, blocking TLR TIR dimerization may ameliorate TLR2-mediated hyperinflammatory states. The BB loop within the TLR TIR domain is critical for mediating certain protein–protein interactions. Examination of the human TLR2 TIR domain crystal structure revealed a pocket adjacent to the highly conserved P681 and G682 BB loop residues. Using computer-aided drug design (CADD), we sought to identify a small molecule inhibitor(s) that would fit within this pocket and potentially disrupt TLR2 signaling. In silico screening identified 149 compounds and 20 US Food and Drug Administration-approved drugs based on their predicted ability to bind in the BB loop pocket. These compounds were screened in HEK293T-TLR2 transfectants for the ability to inhibit TLR2-mediated IL-8 mRNA. C16H15NO4 (C29) was identified as a potential TLR2 inhibitor. C29, and its derivative, ortho-vanillin (o-vanillin), inhibited TLR2/1 and TLR2/6 signaling induced by synthetic and bacterial TLR2 agonists in human HEK-TLR2 and THP-1 cells, but only TLR2/1 signaling in murine macrophages. C29 failed to inhibit signaling induced by other TLR agonists and TNF-α. Mutagenesis of BB loop pocket residues revealed an indispensable role for TLR2/1, but not TLR2/6, signaling, suggesting divergent roles. Mice treated with o-vanillin exhibited reduced TLR2-induced inflammation. Our data provide proof of principle that targeting the BB loop pocket is an effective approach for identification of TLR2 signaling inhibitors.Toll-like receptors (TLRs) are type I transmembrane receptors that detect conserved “pathogen-associated molecular patterns” from microbes, as well as host-derived “danger-associated molecular patterns” (1). TLR2 heterodimerizes with TLR6 or TLR1 to recognize diacyl lipopeptides or triacyl lipopeptides, respectively (2, 3), present in gram-positive and gram-negative bacteria (49).Ligand engagement of TLR2/1 or TLR2/6 activates the myeloid differentiation primary response gene 88 (MyD88)-dependent pathway (i.e., nuclear translocation of NF-κB, activation of MAPKs), resulting in production of proinflammatory cytokines (10). Dysregulated TLR2 signaling has been implicated in numerous diseases (e.g., sepsis, atherosclerosis, tumor metastasis, ischemia/reperfusion injury) (1114). Several inhibitors of TLR2 signaling have been developed (1518), yet none is licensed for human use. A better understanding of the Toll/IL-1 receptor resistance (TIR) domain interactions involved in TLR2 signaling could lead to novel therapeutic agents.Both TLRs and adapter proteins contain a cytoplasmic TIR domain that mediates homotypic and heterotypic interactions during TLR signaling (19). Two adapter proteins implicated in TLR2 signaling are MyD88 and TIRAP (Mal). A conserved Pro [e.g., P681 in human TLR2 (hTLR2), P712 in murine TLR4 (mTLR4), P674 in hTLR10, P804 in mTLR11] within the BB loop of almost all TIR domains is critical for signaling (2027). More importantly, the BB loop P681H mutation in hTLR2 abolished recruitment of MyD88 and signaling (20, 26). Based on this evidence, the BB loop within the TLR2 TIR domain appears to be an ideal target for attenuation of TLR2 signaling.Visual inspection of the crystal structure of the hTLR2 TIR domain (26) revealed a pocket formed by residues on the β-B strand and α-B helix that includes the highly conserved Pro and Gly residues of the BB loop. We hypothesized that targeting this pocket with a small molecule might inhibit interaction of TLR2 with MyD88, and thereby blunt TLR2 signaling. We identified C16H15NO4 (C29) and its derivative, ortho-vanillin (o-vanillin), which inhibit mTLR2 and hTLR2 signaling initiated by synthetic and bacterial agonists without cytotoxicity. Interestingly, mutation of the BB loop pocket residues revealed a differential requirement for TLR2/1 vs. TLR2/6 signaling. Our data indicate that computer-aided drug design (CADD) is an effective approach for identifying small molecule inhibitors of TLR2 signaling and has the potential to identify inhibitors for other TLR signaling pathways.  相似文献   

4.
5.
6.
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   

7.
Cytosolic lipopolysaccharides (LPSs) bind directly to caspase-4/5/11 through their lipid A moiety, inducing inflammatory caspase oligomerization and activation, which is identified as the noncanonical inflammasome pathway. Galectins, β-galactoside–binding proteins, bind to various gram-negative bacterial LPS, which display β-galactoside–containing polysaccharide chains. Galectins are mainly present intracellularly, but their interactions with cytosolic microbial glycans have not been investigated. We report that in cell-free systems, galectin-3 augments the LPS-induced assembly of caspase-4/11 oligomers, leading to increased caspase-4/11 activation. Its carboxyl-terminal carbohydrate-recognition domain is essential for this effect, and its N-terminal domain, which contributes to the self-association property of the protein, is also critical, suggesting that this promoting effect is dependent on the functional multivalency of galectin-3. Moreover, galectin-3 enhances intracellular LPS-induced caspase-4/11 oligomerization and activation, as well as gasdermin D cleavage in human embryonic kidney (HEK) 293T cells, and it additionally promotes interleukin-1β production and pyroptotic death in macrophages. Galectin-3 also promotes caspase-11 activation and gasdermin D cleavage in macrophages treated with outer membrane vesicles, which are known to be taken up by cells and release LPSs into the cytosol. Coimmunoprecipitation confirmed that galectin-3 associates with caspase-11 after intracellular delivery of LPSs. Immunofluorescence staining revealed colocalization of LPSs, galectin-3, and caspase-11 independent of host N-glycans. Thus, we conclude that galectin-3 amplifies caspase-4/11 oligomerization and activation through LPS glycan binding, resulting in more intense pyroptosis—a critical mechanism of host resistance against bacterial infection that may provide opportunities for new therapeutic interventions.

Lipopolysaccharides (LPSs) are pathogen-associated molecular patterns that can elicit a host defense response through binding to cell-surface Toll-like receptor 4 (TLR4). Systemic inflammatory response syndrome is induced by overstimulation of the innate immune response via LPSs, resulting in severe multiple organ failure, which is a major cause of death worldwide in intensive care units (1). LPS-induced dimerization of TLR4 initiates signal transduction involving the NF-κB– and MyD88-dependent and -independent pathways, thereby contributing to various inflammatory responses (2). Another set of the immune repertoire, which resides in the cytosol and comprises NLRP1, NLRP3, NAIP/NLRC4, and AIM2, is known as the inflammasome. Inflammasomes can be activated in response to a number of well-defined pathogen-derived ligands and physiological aberrations, which in turn trigger caspase-1–mediated pyroptotic death (3, 4). This process has been associated with strengthening the host defense program to eliminate intracellular bacteria.Recently, a cytosolic LPS-sensing pathway involving caspase-4/5 in humans and caspase-11 in mice was termed the noncanonical inflammasome pathway, and this pathway is independent of TLR4 (58). LPSs from extracellular bacteria can enter the cytoplasm and trigger caspase-4/5/11–dependent responses. LPSs can be delivered into the cytosol when LPS-containing outer membrane vesicles (OMVs) from gram-negative bacteria are taken up by the cells or when intracellular bacteria escape from the phagosomes that are damaged by host resistant factors such as guanylate-binding protein and HMGB1 or microbe-derived hemolysins (912). LPSs comprise three regions: lipid A, core oligosaccharide, and O-polysaccharide (also termed O-antigen). The lipid A moiety binds directly to the caspase-4/5/11 caspase activation and recruitment domain (CARD, also known as prodomain), leading to caspase oligomerization and activation (7). This event likely mimics the proximity-induced dimerization model of initiator caspase activation (13). Furthermore, caspase-4/5/11 executes downstream signaling events via gasdermin D. Activated inflammatory caspase proteolytically cleaves gasdermin D to create an N-terminal fragment that self-oligomerizes and then inserts into the cell membrane to form pores, causing lytic cell death (1417). Various stimuli have been identified in the caspase-1–mediated canonical-inflammasome signaling pathway (3, 4), but the detailed mechanism underlying noncanonical inflammasome activation mediated by caspase-4/5/11 remains unclear.Galectins, a family of β-galactoside–binding proteins, can decode host-derived complex glycans and are involved in various biological responses (1823). Galectins are nucleocytoplasmic proteins synthesized without a classical signal sequence, although they can be secreted through unconventional pathways (19, 21, 23, 24). Recent studies have revealed prominent roles of cytosolic galectins in host defense programs (12, 25, 26). The proposed molecular mechanisms involve the binding of galectins to host glycans exposed to the cytosolic milieu upon endosomal or phagosomal membrane damage. In addition to binding host glycans, galectins also recognize microbial glycans, particularly LPSs (2730). However, the contribution of galectins to the host response through binding to cytosolic LPSs is unknown.Galectin-3 is an ∼30-kDa protein that contains a carbohydrate-recognition domain (CRD) connected to N-terminal proline, glycine, and tyrosine-rich tandem repeats. Upon binding to multivalent glycoconjugates through its CRD, the protein forms oligomers, which is attributable to the self-association property of its N-terminal region (31, 32). Galectin-3 binds to LPSs of various gram-negative bacteria by recognizing their carbohydrate residues (3336).Although structural information is scarce (37), existing information suggests that ligand-induced oligomerization of caspase CARD is necessary for the activation of inflammatory caspases (7, 38). Therefore, we hypothesized that galectin-3 may be an intracellular LPS sensor that participates in LPS-induced CARD-mediated inflammatory caspase activation. Specifically, highly ordered arrays of LPS–galectin-3 complexes may amplify caspase-4/5/11 oligomerization and activation. Here, we investigated the formation of galectin-3–LPS–caspase-4/11 complexes in cell-based and cell-free systems. Our findings provide evidence regarding a role of galectin-3 as an intracellular mediator in noncanonical inflammasome activation through LPS glycan recognition.  相似文献   

8.
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   

9.
Microglia (tissue-resident macrophages) represent the main cell type of the innate immune system in the CNS; however, the mechanisms that control the activation of microglia are widely unknown. We systematically explored microglial activation and functional microglia–neuron interactions in organotypic hippocampal slice cultures, i.e., postnatal cortical tissue that lacks adaptive immunity. We applied electrophysiological recordings of local field potential and extracellular K+ concentration, immunohistochemistry, design-based stereology, morphometry, Sholl analysis, and biochemical analyses. We show that chronic activation with either bacterial lipopolysaccharide through Toll-like receptor 4 (TLR4) or leukocyte cytokine IFN-γ induces reactive phenotypes in microglia associated with morphological changes, population expansion, CD11b and CD68 up-regulation, and proinflammatory cytokine (IL-1β, TNF-α, IL-6) and nitric oxide (NO) release. Notably, these reactive phenotypes only moderately alter intrinsic neuronal excitability and gamma oscillations (30–100 Hz), which emerge from precise synaptic communication of glutamatergic pyramidal cells and fast-spiking, parvalbumin-positive GABAergic interneurons, in local hippocampal networks. Short-term synaptic plasticity and extracellular potassium homeostasis during neural excitation, also reflecting astrocyte function, are unaffected. In contrast, the coactivation of TLR4 and IFN-γ receptors results in neuronal dysfunction and death, caused mainly by enhanced microglial inducible nitric oxide synthase (iNOS) expression and NO release, because iNOS inhibition is neuroprotective. Thus, activation of TLR4 in microglia in situ requires concomitant IFN-γ receptor signaling from peripheral immune cells, such as T helper type 1 and natural killer cells, to unleash neurotoxicity and inflammation-induced neurodegeneration. Our findings provide crucial mechanistic insight into the complex process of microglia activation, with relevance to several neurologic and psychiatric disorders.Microglia are tissue-resident macrophages in the CNS that become activated in most brain disorders, such as bacterial meningoencephalitis, multiple sclerosis, and Alzheimer’s disease (1, 2). Activation of microglia features changes in morphology and receptor expression, antigen presentation, cytokine release, migration, and phagocytosis, and it ranges from proinflammatory and potentially neurotoxic to anti-inflammatory and neuroprotective phenotypes (1, 3, 4). The mechanisms that control the transition of microglia to reactive phenotypes, including the impact on neuronal function, are mostly unknown, however (57).Sensing of microbial or modified endogenous ligands by microglia is mediated by innate pattern recognition receptors, such as scavenger receptors and Toll-like receptors (TLRs). A prime example is TLR4, which acts with CD14, MD-2, and lipopolysaccharide (LPS)-binding protein in recognizing LPS, a cell wall component of Gram-negative bacteria (8, 9). TLR4 is also central to microglial recognition of amyloid-β peptide, which is thought to be part of the inflammatory response in Alzheimer’s disease (7, 10).LPS has been widely used to study the molecular mechanisms of microglial activation in inflammatory neurodegeneration (13). In primary monocultures and microglia-neuron cultures, LPS exposure alone or in combination with IFN-γ for a “booster” triggers the massive release of proinflammatory and cytotoxic factors, such as TNF-α, IL-6, and nitric oxide (NO), finally resulting in neuronal death (8, 1118). Similar effects were observed in vivo after intracerebral administration of LPS (1921). These and other studies have contributed to the concept that microglial TLR4 activation with LPS (i.e., with a single pathogenic stimulus) is sufficient to induce neurodegeneration (22, 23); however, this concept is biologically risky, and has been questioned in some experimental works and reviews (24, 11, 24, 25).Most previous studies focused on two aspects of microglial TLR4 activation with LPS: (i) the properties of the reactive microglial phenotype(s) and (ii) the degree of neurodegeneration. For this purpose, either simple culture systems or in vivo models, in which interactions with leukocytes infiltrating from the blood are inevitable, have been used (1, 4). Thus, it is widely unknown how TLR4 and IFN-γ receptor signaling in microglia individually contribute to neurotoxicity and neurodegeneration in situ. This aspect is highly relevant for several neurologic and psychiatric disorders. Moreover, concomitant alterations in neuronal information processing (i.e., dysfunction in excitatory pyramidal cells and inhibitory GABAergic interneurons, including astrocytes) have been little explored (2527).We rigorously addressed these fundamental questions in postnatal neuronal tissue (1, 4). To mimic microglial confrontation with LPS in situ and, notably, in the absence of infiltrating leukocytes, we used organotypic hippocampal slice cultures that feature highly preserved cytoarchitectures and complex neuronal network functions (5, 28). Microglial interaction with infiltrating T helper type 1 (Th1) cells and/or natural killer (NK) cells was mimicked by recombinant IFN-γ administration.  相似文献   

10.
Horizontal gene transfer (HGT) provides an evolutionary shortcut for recipient organisms to gain novel functions. Although reports of HGT in higher eukaryotes are rapidly accumulating, in most cases the evolutionary trajectory, metabolic integration, and ecological relevance of acquired genes remain unclear. Plant cell wall degradation by HGT-derived enzymes is widespread in herbivorous insect lineages. Pectin is an abundant polysaccharide in the walls of growing parts of plants. We investigated the significance of horizontally acquired pectin-digesting polygalacturonases (PGs) of the leaf beetle Phaedon cochleariae. Using a CRISPR/Cas9-guided gene knockout approach, we generated a triple knockout and a quadruple PG-null mutant in order to investigate the enzymatic, biological, and ecological effects. We found that pectin-digestion 1) is exclusively linked to the horizontally acquired PGs from fungi, 2) became fixed in the host genome by gene duplication leading to functional redundancy, 3) compensates for nutrient-poor diet by making the nutritious cell contents more accessible, and 4) facilitates the beetles development and survival. Our analysis highlights the selective advantage PGs provide to herbivorous insects and demonstrate the impact of HGT on the evolutionary success of leaf-feeding beetles, major contributors to species diversity.

Based on a solid fossil record of feeding damage starting from the Devonian, plants have been eaten by insects for more than 400 My (1). Nowadays, both vascular plants and herbivorous insects constitute most of the macroscopic biological diversity (24). Herbivorous insects—which make up a quarter of all eukaryote species—outnumber every other plant-consuming clade of animals (5). This leads to a fundamental question in biology: which insect traits promoted their evolutionary success? Insect species richness has often been attributed to the diversity of their hostplants, a controversial correlation that depends on the specific clade and taxonomic level analyzed (69). Although the major evolutionary transition toward a plant-based diet remains elusive, the long-lasting interaction of insects with their hosts resulted in a plethora of morphological, physiological, and behavioral adaptations (6). Consequently, a better understanding of these adaptations would shed light on fundamental mechanisms by which herbivorous insects successfully radiated and persisted on plants.Besides the well-known chemical defenses of plants against pathogens and herbivores (10), structural defenses like the plant cell wall (PCW) make it difficult for an insect to access essential nutrients (11).Variations in PCW density and composition affect food choice and performance of herbivorous insects (12, 13). In turn, survival on a fiber-rich but nutrient-poor diet requires adaptations such as strong mandibles, compensatory feeding, or associations with nutritional symbionts (1416).The PCW is difficult to digest because it is a composite material made of interconnected polysaccharides (17, 18). The primary wall is dominant in growing plant tissues and consists of a network of cellulose and hemicelluloses embedded in a pectin matrix (19). Pectin is a hetero-polysaccharide with a backbone rich in galacturonic acid (GalUA) and is highly abundant in primary walls and in the cell-connecting middle lamellae (20). Besides its overall protective role in contributing to wall strength and cell–cell adhesion, pectin also mediates defense against microbial pathogens (21). To weaken the PCW, fungal and bacterial pathogens secrete an array of PCW degrading enzymes (PCWDEs), including enzymes that digest pectin, important virulence factors (2224).Pectinolytic enzymes, as an adaptation to access plant nutrients, were long thought to be restricted to bacteria and fungi (25). Insect pectinase activity and other PCWDEs were believed to be restricted to symbiotic microbes associated with their hosts (26, 27). This assumption was first challenged by the discovery of a termite-endogenous cellulase (28) and the finding that a pectinolytic polygalacturonase (PG) was encoded in the genome of the weevil Sitophilus oryzae (29). Since then, an increasing number of animal-endogenous PGs were identified in the genomes of Phasmatodea (stick insects) (30), Hemiptera (true bugs) (31), and Coleoptera (beetles) (3236). Recent studies have shown that insect-derived PGs, belonging to family 28 of glycoside hydrolases (GH28) cleave the GalUA backbone either terminally (exo-PG) or randomly (endo-PG), similar to their microbial counterparts (3739).A fascinating aspect of insect PGs is their evolutionary origin by horizontal gene transfer (HGT). Phylogenetic analyses have suggested several independent events of microbe-to-insect HGT, consistent with the patchy distribution of PGs in herbivores (3941). Their conservation for millions of years after initial acquisition, metabolic integration, and subfunctionalization following gene duplication suggest that PGs played a key role in the diversification of plant-feeding insects (38, 41). The positive correlation between species richness and the appearance of PCWDEs including PGs has indeed been shown recently in the Phytophaga (42), a superradiation comprising weevils, leaf beetles, and longhorned beetles (43, 44).Despite these correlative studies, experimental evidence of the adaptive value of insect PGs is still lacking. We had previously established that the leaf beetle Phaedon cochleariae, possessed genes for three enzymatically active endo-PGs (GH28-1, GH28–5, and GH28–9) which were specifically expressed in the beetle’s gut (38, 45). Further sequencing revealed a fourth PG gene (GH28-10) that was previously considered a GH28-9 allele due to high sequence similarities. We sequentially knocked out all four PG-encoding genes, to assess the importance of pectin digestion on larval growth, developmental time, and survival. We found that a decrease in PG activity correlates with a lower fitness in mutant individuals, due to their inability to efficiently digest pectin. We rescued the fitness of the PG knockout mutants by supplementing their diet with a recombinant form of their own GH28-1 enzyme. In contrast, these mutants could not be rescued by adding a pectin breakdown product in their diet. This indicates that breaking open the PCW to obtain nutrient-rich cytoplasm is the primary benefit of pectinases rather than the production of pectin digestion projects for nutrition. This is an experimental demonstration of the importance of PCW digestion to herbivore fitness. These results suggest that pectinases acquired by HGT, and PCWDEs in general, represent key innovations, in part explaining the evolutionary success of herbivorous beetles.  相似文献   

11.
Escherichia coli infections, a leading cause of septic shock, remain a major threat to human health because of the fatal action to endotoxin (LPS). Therapeutic attempts to neutralize endotoxin currently focus on inhibiting the interaction of the toxic component lipid A with myeloid differentiating factor 2, which forms a trimeric complex together with Toll-like receptor 4 to induce immune cell activation. The 1.73-Å resolution structure of the unique endotoxin-neutralizing protective antibody WN1 222-5 in complex with the core region shows that it recognizes LPS of all E. coli serovars in a manner similar to Toll-like receptor 4, revealing that protection can be achieved by targeting the inner core of LPS and that recognition of lipid A is not required. Such interference with Toll-like receptor complex formation opens new paths for antibody sepsis therapy independent of lipid A antagonists.LPS from Gram-negative bacteria is the major etiological agent of septic shock, which is a serious and often fatal dysregulation of the innate immune response that affects 750,000 people in the United States annually (1). Infection with Escherichia coli, together with Klebsiella, Neisseria, and Pseudomonas, are the most frequent isolates in septic shock (2). A key event initiating the shock cascade is the induction of the innate immune response by the complex formation of a symmetric “m”-shaped multimer composed of two copies of Toll-like receptor 4 (TLR4), myeloid differentiating factor 2 (MD-2), and LPS (3, 4). In a landmark publication, the structure of TLR4-MD-2 bound to LPS (3) was recently described.LPS is composed of an acylated glucosamine phosphate disaccharide (i.e., lipid A), which is the endotoxic principle of LPS, a core oligosaccharide (core-OS) and a distal O-polysaccharide (O-PS) often composed of repeating units (Fig. 1A). Whereas the O-PS is structurally heterogeneous, with more than 180 reported E. coli serotypes (5), the core region is composed of a more conserved structure commonly divided into the inner Kdo-heptose and outer hexose regions (6).Open in a separate windowFig. 1.Structures of LPS and the shape of the combining site. (A) Structure of E. coli R2 dodecasaccharide-P4, representing the core and lipid A of the LPS from Enterobacteria commonly associated with septic shock. (B) Stereo views of electron density corresponding to 10 sugar residues of the core antigen (the lipid A moiety is disordered) contoured at 1.0 σ.Recognition of LPS leads to a paramount immunological defense reaction caused by the activation of a complex network of immunological mediators. Attempts to control the clinical development of sepsis by neutralizing the most important proinflammatory mediators have failed, including the recent withdrawal of recombinant activated protein C (Xigris). A promising antagonistic lipid candidate called Eritoran (E5564; Eisai) (7) also recently failed in clinical trials, and alternative treatments are urgently needed. The discovery of TLR4 as the principal receptor for endotoxins (8) has stimulated the development of drugs aiming at its down-regulation (9) through interference of LPS–TLR4–MD-2 complex formation (4, 1012).Antisera specific for O-PS have been shown to protect against LPS lethality (13); however, the diversity of enterobacterial O-PS together with the rapid onset of septic shock have hindered their introduction into clinical practice (11).The hypothesis that mAbs specific to the conserved inner core region or lipid A would be protective against a wide range of serovars and even different species was put forward (14) after the discovery of structural similarities within their respective LPSs. WN1 222-5 is the only neutralizing antibody reported to date that displays specificity for an epitope within the structurally conserved region of LPS from a large number of pathogenic E. coli, Salmonella, Shigella, and Citrobacter serovars (15). Further, WN1 222-5 has been shown to inhibit the recognition and uptake of LPS by cells expressing coreceptor mCD14, likely by hindering the transfer of LPS to TLR4–MD-2 (16).WN1 222-5 has been shown to inhibit the inflammatory cascade in in vivo studies of septic shock, in which it prevents the pyrogenic response in rabbits, inhibits the Limulus amoebocyte lysate assay, and inhibits LPS-induced monokine secretion (1517).The difficulties in growing crystals of antibodies in complex with carbohydrate antigens has led to relatively few reported structures (1821), leading, for example, to increased use of structure prediction tools such as molecular dynamics modeling (22). Thus, in contrast to their great immunological significance during infectious disease, still relatively little is known about carbohydrate recognition by antibodies at the structural level. Whereas cavity- or groove-shaped antibody-combining sites have been observed in most cases, a unique mechanism of binding has been observed for the HIV-1 neutralizing antibody 2G12, binding clusters of carbohydrates from the silent face of gp120 by using “domain swapping” (19, 23, 24).The structural analysis of antibodies Se155-4 and S20-4 against O-PS of Salmonella enterica and Vibrio cholerae, respectively, have revealed structural insights into the high specificity for a particular serotype (20, 25). However, because of their specificity, antibodies against O-PS are of limited use for the treatment of infectious disease. Nevertheless, structures of antibodies in complex with large carbohydrate antigens have revealed critical insights for vaccine development. The protective antibody F22-4 in complex with an 11-sugar segment from the O-PS of Shigella flexneri serotype 2a (26) allowed the design of new immunogens.Most attempts in obtaining antibodies that are broadly reactive with a wide variety of LPSs from different Gram-negative bacteria have failed, and epitopes within the deeper core region of LPS have been regarded as not accessible to antibodies in WT LPSs of infectious bacteria. To provide detailed insight on a unique cross-reactive and neutralizing ability, the Fab from WN1 222-5 in complex with a complete core-OS of LPS from E. coli has been crystallized and its structure determined to 1.73-Å resolution.  相似文献   

12.
Wild and managed bees are well documented as effective pollinators of global crops of economic importance. However, the contributions by pollinators other than bees have been little explored despite their potential to contribute to crop production and stability in the face of environmental change. Non-bee pollinators include flies, beetles, moths, butterflies, wasps, ants, birds, and bats, among others. Here we focus on non-bee insects and synthesize 39 field studies from five continents that directly measured the crop pollination services provided by non-bees, honey bees, and other bees to compare the relative contributions of these taxa. Non-bees performed 25–50% of the total number of flower visits. Although non-bees were less effective pollinators than bees per flower visit, they made more visits; thus these two factors compensated for each other, resulting in pollination services rendered by non-bees that were similar to those provided by bees. In the subset of studies that measured fruit set, fruit set increased with non-bee insect visits independently of bee visitation rates, indicating that non-bee insects provide a unique benefit that is not provided by bees. We also show that non-bee insects are not as reliant as bees on the presence of remnant natural or seminatural habitat in the surrounding landscape. These results strongly suggest that non-bee insect pollinators play a significant role in global crop production and respond differently than bees to landscape structure, probably making their crop pollination services more robust to changes in land use. Non-bee insects provide a valuable service and provide potential insurance against bee population declines.Pollinator-dependent crops are increasingly grown to provide food, fiber, and fuel as well as micronutrients essential to human health (15). The yield and quality of these crops benefit to varying degrees from flower visitation by animals. The honey bee, Apis mellifera L. (Hymenoptera: Apidae), is the most versatile, ubiquitous, and commonly used managed pollinator (6), but the global reliance on this single pollinator species is a risky strategy, especially given major threats to the health of managed honey bee colonies because of poor nutrition, the ectoparasitic mite Varroa destructor Anderson and Trueman (Mesostigmata: Varroidae), and a number of other pests and diseases (710).However, honey bees are not the only insects that pollinate crops. Apart from a few managed bee taxa, the great majority of other pollinators are free-living or wild, providing an ecosystem service to crops. Wild pollinators other than honey bees recently have been recognized for their role in increasing and stabilizing crop-pollination services (11, 12). Wild bees are known to improve seed set, quality, shelf life, and commercial value of a variety of crops (1317). Increasingly, studies indicate that insect pollinators other than bees, such as flies, beetles, moths, and butterflies, are equally if not more important for the production of some crops (1824). Nonetheless, the contribution to crop pollination by non-bee insects has been largely unnoticed, with most global syntheses focusing on bees (2528) or grouping together all bee and non-bee wild-insect pollinators (11).Diverse pollinator assemblages have been shown to increase pollination services as a result of complementary resource use arising from variations in morphology and behavior among pollinator taxa (29, 30). For example, pollinator species may visit different parts within a flower or inflorescence or different flowers within a plant (high versus low flowers), improving the quality or quantity of pollination services overall (13, 3133). Non-bee taxa, in particular, often have broader temporal activity ranges (3436) and can provide pollination services at different times of the day compared with bees and in weather conditions when bees are unable to forage (3740). In addition, non-bee taxa may be more efficient in transferring pollen for some crops under certain conditions (18, 19, 38) and/or carry pollen further distances than some bees (41). It has been suggested that this long-distance pollen transfer could have important genetic consequences for wild plants (42, 43). However, there is little information on the overall importance of the diverse group of non-bee wild pollinators (but see refs. 39 and 44) and their importance to global crop production.Anthropogenic land use change and intensification are considered to be among the main drivers of bee declines (45, 46). One of the mechanisms underlying observed declines is thought to be the loss of habitat that supports host plants (47) and nesting sites (48). However, different pollinator taxa respond differently to disturbances (49, 50). The proximity and area of natural habitat are often associated with higher crop flower visitation and bee diversity (25, 46, 51). Yet, although several studies have investigated the habitat requirements of non-bee taxa (5255), little is known about how habitat availability affects crop-pollination services from non-bee taxa (but see ref. 44). Thus, differential responses to habitat proximity by bees and non-bees, if such exists, could provide an additional stabilizing effect on crop-pollination services.In summary, non-bees are often neglected as potential providers of crop ecosystem services by the scientific community and by growers. In the data collection for the present synthesis, for example, 33% of the original 58 pollination studies we obtained did not record or distinguish non-bee pollinators from bee pollinators and thus had to be excluded.In this study we address the knowledge gap about non-bee crop pollination and ask:
  • i)How does the crop pollination provided by non-bee insects compare with that provided by honey bees and other bees?
  • ii)How does the crop pollination provided by non-bees, honey bees, and other bees translate into fruit/seed set?
  • iii)Do non-bee crop pollinators respond similarly to bees with regard to isolation from natural and semi/natural habitats?
To answer these questions, we compiled a dataset comprising 39 studies of crop pollinators around the world and the pollination services they provide (Table S1).  相似文献   

13.
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.
Aberrant immune responses, including hyperresponsiveness to Toll-like receptor (TLR) ligands, underlie acute respiratory distress syndrome (ARDS). Type I interferons confer antiviral activities and could also regulate the inflammatory response, whereas little is known about their actions to resolve aberrant inflammation. Here we report that interferon-β (IFN-β) exerts partially overlapping, but also cooperative actions with aspirin-triggered 15-epi-lipoxin A4 (15-epi-LXA4) and 17-epi-resolvin D1 to counter TLR9-generated cues to regulate neutrophil apoptosis and phagocytosis in human neutrophils. In mice, TLR9 activation impairs bacterial clearance, prolongs Escherichia coli–evoked lung injury, and suppresses production of IFN-β and the proresolving lipid mediators 15-epi-LXA4 and resolvin D1 (RvD1) in the lung. Neutralization of endogenous IFN-β delays pulmonary clearance of E. coli and aggravates mucosal injury. Conversely, treatment of mice with IFN-β accelerates clearance of bacteria, restores neutrophil phagocytosis, promotes neutrophil apoptosis and efferocytosis, and accelerates resolution of airway inflammation with concomitant increases in 15-epi-LXA4 and RvD1 production in the lungs. Pharmacological blockade of the lipoxin receptor ALX/FPR2 partially prevents IFN-β–mediated resolution. These findings point to a pivotal role of IFN-β in orchestrating timely resolution of neutrophil and TLR9 activation–driven airway inflammation and uncover an IFN-β–initiated resolution program, activation of an ALX/FPR2-centered, proresolving lipids-mediated circuit, for ARDS.

Acute respiratory distress syndrome (ARDS) is a common syndrome associated with high mortality in patients admitted to intensive care units (1). ARDS is characterized by diffuse alveolar damage that develops in patients with known risk factors, most commonly pneumonia, sepsis, or trauma (2, 3). The initial alveolar damage leads to recruitment of neutrophils and monocytes, which further aggravate injury (3). Treatment of the underlying cause and lung-protective ventilation are the main elements of supportive therapy (4, 5). Importantly, no therapies are available to resolve the aberrant immune responses underlying ARDS.Type I interferons, IFN-α and IFN-β, are well established to confer antiviral activities to host cells and could also regulate the inflammatory response. A delayed type I interferon response triggers the generation of proinflammatory cytokines and facilitates the recruitment of monocytes to the lung, resulting in lethal pneumonia in mice infected with SARS-CoV-1 (6) or SARS-CoV-2 (7). Type I interferons break TNF-induced tolerance to Toll-like receptor (TLR) signals on monocytes/macrophages, rendering them hyperresponsive to additional TLR signals concurrent with inflammatory activation (8). For instance, bacterial DNA (CpG DNA) or mitochondrial DNA through TLR9 impairs neutrophil phagocytosis, delays neutrophil apoptosis, and perpetuates inflammation (9, 10). In contrast, IFN-β protects against lethal polymicrobial sepsis through inhibiting IL-1 production and/or induction of IL-10 (1113). IFN-β produced by macrophages during resolution of bacterial pneumonia facilitates removal of neutrophils from inflamed tissues and reprograms macrophages to a proresolving phenotype, thereby driving inflammatory resolution in mice (14). However, the underlying mechanisms are incompletely understood; albeit these would be essential for implementing precision treatment with IFN-β.Resolution of inflammation is an active process governed by specialized proresolving lipid and protein mediators (SPMs) (1519). These mediators converge on select receptors, including the pleiotropic lipoxin A4 receptor/formyl peptide receptor 2 (ALX/FPR2) (20). ALX/FPR2 plays critical roles in host defense and orchestrating inflammatory resolution (2023). ALX/FPR2 binds multiple lipid ligands, including aspirin-triggered 15-epi-lipoxin A4 (15-epi-LXA4) and 17-epi-resolvin D1 (17-epi-RvD1), generated within the inflammatory microenvironment (17, 18). SPMs inhibit neutrophil recruitment, promote neutrophil apoptosis and efferocytosis, and facilitate tissue repair and return to homeostasis (17, 18, 24). Activation of ALX/FPR2 with 15-epi-LXA4 or 17-epi-RvD1 counters TLR9-generated cues, restores impaired neutrophil function, and enhances timely resolution of airway bacterial infections (9). Since resolution of inflammation is skewed toward a proresolving lipid profile (18, 25, 26), we investigated whether IFN-β can modulate ALX/FPR2-based resolution mechanisms. Here, we report that IFN-β exerts partially overlapping, but also cooperative actions with 17-epi-RvD1 to counter TLR9-generated signals to regulate neutrophil phagocytosis and apoptosis in vitro. In mice, IFN-β facilitates clearance of bacteria, neutrophil apoptosis and efferocytosis, and promotes the resolution of acute airway inflammation, in part, by stimulating generation of proresolving lipids and activation of ALX/FPR2-centered proresolving circuits. Our results uncover a hitherto unrecognized effector mechanism by which IFN-β may facilitate resolution of ARDS.  相似文献   

17.
The protumor roles of alternatively activated (M2) tumor-associated macrophages (TAMs) have been well established, and macrophage reprogramming is an important therapeutic goal. However, the mechanisms of TAM polarization remain incompletely understood, and effective strategies for macrophage targeting are lacking. Here, we show that miR-182 in macrophages mediates tumor-induced M2 polarization and can be targeted for therapeutic macrophage reprogramming. Constitutive miR-182 knockout in host mice and conditional knockout in macrophages impair M2-like TAMs and breast tumor development. Targeted depletion of macrophages in mice blocks the effect of miR-182 deficiency in tumor progression while reconstitution of miR-182-expressing macrophages promotes tumor growth. Mechanistically, cancer cells induce miR-182 expression in macrophages by TGFβ signaling, and miR-182 directly suppresses TLR4, leading to NFκb inactivation and M2 polarization of TAMs. Importantly, therapeutic delivery of antagomiR-182 with cationized mannan-modified extracellular vesicles effectively targets macrophages, leading to miR-182 inhibition, macrophage reprogramming, and tumor suppression in multiple breast cancer models of mice. Overall, our findings reveal a crucial TGFβ/miR-182/TLR4 axis for TAM polarization and provide rationale for RNA-based therapeutics of TAM targeting in cancer.

It is well known that the nonmalignant stromal components in tumors play pivotal roles in tumor progression and therapeutic responses (1, 2). Macrophages are a major component of tumor microenvironment and display considerable phenotypic plasticity in their effects toward tumor progression (35). Classically activated (M1) macrophages often exert direct tumor cytotoxic effects or induce antitumor immune responses by helping present tumor-related antigens (6, 7). In contrast, tumoral cues can polarize macrophages toward alternative activation with immunosuppressive M2 properties (68). Numerous studies have firmly established the protumor effects of M2-like tumor-associated macrophages (TAMs) and the association of TAMs with poor prognosis of human cancer (911). However, how tumors induce the coordinated molecular and phenotypic changes in TAMs for M2 polarization remains incompletely understood, impeding the designing of TAM-targeting strategies for cancer intervention. In addition, drug delivery also represents a hurdle for therapeutic macrophage reprogramming.Noncoding RNAs, including microRNAs, have been shown to play vital roles in various pathological processes of cancer (12). The microRNA miR-182 has been implicated in various developmental processes and disease conditions (1315). Particularly, it receives extensive attention in cancer studies. Prevalent chromosomal amplification of miR-182 locus and up-regulation of its expression in tumors have been observed in numerous cancer types including breast cancer, gastric cancer, lung adenocarcinoma, colorectal adenocarcinoma, ovarian carcinoma, and melanoma (1621). miR-182 expression is also linked to higher risk of metastasis and shorter survival of patients (20, 2224). Functional studies showed that miR-182 expression in cancer cells plays vital roles in various aspects of cancer malignancy, including tumor proliferation (2529), migration (30, 31), invasion (16, 32, 33), epithelial-mesenchymal transition (3436), metastasis (21, 37, 38), stemness (30, 39, 40), and therapy resistance (41, 42). A number of target genes, including FOXO1/3 (18, 21, 4345), CYLD (46), CADM1 (47), BRCA1 (27, 48), MTSS1 (34), PDK4 (49), and SMAD7 (35), were reported to be suppressed by miR-182 in cancer cells. Our previous work also proved that tumoral miR-182 regulates lipogenesis in lung adenocarcinoma and promotes metastasis of breast cancer (34, 35, 49). Although miR-182 was established as an important regulator of cancer cell malignancy, previous studies were limited, with analyses of miR-182 in cultured cancer cells and transplanted tumors. Thus, the consequences of miR-182 regulation in physiologically relevant tumor models, such as genetically modified mice, have not been shown. More importantly, whether miR-182 also plays a role in tumor microenvironmental cell components is unknown.In this study, we show that miR-182 expression in macrophages can be induced by breast cancer cells and regulates TAM polarization in various tumor models of mice. In addition, miR-182 inhibition with TAM-targeting exosomes demonstrates promising efficacy for cancer treatment.  相似文献   

18.
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   

19.
CpG-oligodeoxynucleotides (CpG-ODNs) are potent immune stimuli currently under investigation as antimicrobial agents for different species. Toll-like receptor (TLR) 9 and TLR21 are the cellular receptors of CpG-ODN in mammals and chickens, respectively. The avian genomes lack TLR9, whereas mammalian genomes lack TLR21. Although fish contain both of these genes, the biological functions of fish TLR9 and TLR21 have not been investigated previously. In this study, we comparatively investigated zebrafish TLR9 (zebTLR9) and TLR21 (zebTLR21). The two TLRs have similar expression profiles in zebrafish. They are expressed during early development stages and are preferentially expressed in innate immune function-related organs in adult fish. Results from cell-based activation assays indicate that these two zebrafish TLRs are functional, responding to CpG-ODN but not to other TLR ligands. zebTLR9 broadly recognized CpG-ODN with different CpG motifs, but CpG-ODN with GACGTT or AACGTT had better activity to this TLR. In contrast, zebTLR21 responded preferentially to CpG-ODN with GTCGTT motifs. The distinctive ligand recognition profiles of these two TLRs were determined by their ectodomains. Activation of these two TLRs by CpG-ODN occurred inside the cells and was modulated by UNC93B1. The biological functions of these two TLRs were further investigated. The CpG-ODNs that activate both zebTLR9 and zebTLR21 were more potent than others that activate only zebTLR9 in the activation of cytokine productions and were more bactericidal in zebrafish. These results suggest that zebTLR9 and zebTLR21 cooperatively mediate the antimicrobial activities of CpG-ODN. Overall, this study provides a molecular basis for the activities of CpG-ODN in fish.Bacterial and viral CpG-deoxynucleotides containing DNA (CpG-DNA) represent a type of pathogen-associated molecular pattern (PAMP) that activates immune cells and triggers host responses to microbial infections (13). Synthetic phosphorothioate-modified CpG-oligodeoxynucleotides (CpG-ODNs) mimic the functions of CpG-DNA and have been investigated as immune modulators for their adjuvant and antimicrobial activities in different species (47). In general, a CpG-ODN contains one or more copies of CpG-deoxynucleotides containing hexamer motifs (CpG motifs). A CpG-ODN’s immunostimulatory activities are dependent on its length, the number of CpG motifs, and the position, spacing, and surrounding bases of these CpG motifs.A CpG-ODN can have varying immunostimulatory activity in different species. This species-specific property is determined by the nucleotide context of the CpG motifs within the CpG-ODN. For example, CpG-ODNs containing a purine-purine-CG-pyrimidine-pyrimidine motif, such as a GACGTT motif, are more potent in activating murine cells compared with those containing a GTCGTT motif. In contrast, the GTCGTT motif containing CpG-ODN generates stronger immune responses in humans and various domestic animals (8, 9).Toll-like receptors (TLRs) are pattern recognition receptors that play crucial roles in the initiation of host defense against microbial invasion by binding to PAMPs from the invading microorganisms. Ten TLRs (TLR1–TLR10) have been identified in human cells, and 13 have been identified in mouse cells. These TLRs detect diverse structures of PAMP from lipids, lipoproteins, glycans, and proteins to nucleic acids (10, 11). Of these, TLR9, a member of a subfamily of intracellular TLRs comprising TLR3, TLR7, TLR8, and TLR9, is the cellular receptor that mediates the functions of CpG-ODN. The species-specific activity of a CpG-ODN is attributed to a species-specific ligand recognition of TLR9 (1214). In mammals, cellular localization and activation of TLR9 are regulated by various accessory proteins, including UNC93 Caenorhabditis elegans homolog of B1 (UNC93B1) (1517). Activation of TLR9 by CpG-ODN results in various immunologic effects, including up-regulation of MHC class I and II costimulatory molecules, activation of natural killer cells and B cells, and increased B-cell proliferation. In addition, TLR9 activation up-regulates T helper (Th) 1-polarized cytokine production, which promotes T-cell activation. Because of these potent immunostimulatory effects, CpG-ODNs are currently under investigation for various therapeutic applications, including antitumor and anti-infection therapies and as vaccine adjuvants (1820).Similar to their actions in mammalian species, in chickens CpG-ODNs activate marked immune responses and provide protection from microbial infections (4, 5, 21). Nevertheless, analysis of the chicken and zebra finch genomes found that the TLR9 gene is not present in avian genomes. Of the 10 avian TLRs, TLR1La, TLR1Lb, TLR2a, TLR2b, TLR3, TLR4, TLR5, and TLR7 are orthologs to mammalian TLRs, whereas TLR15 and TLR21 are not found in mammals (22). It was recently demonstrated that chicken TLR21 (chTLR21) is a functional homolog to mammalian TLR9 in terms of response to CpG-ODN stimulation (23, 24).The immunostimulatory effects of CpG-ODNs have been investigated in numerous fish species as well. In these species, much like in mammalian and avian species, CpG-ODNs up-regulate the activation of macrophages, induce proliferation of leukocytes, and stimulate cytokine expression. In addition, CpG-ODNs have been shown to protect fish against bacterial and viral infections. The molecular bases for CpG-ODN activation in fish remain unclear, however (5, 6). The genomic DNA of zebrafish has been sequenced and annotated, leading to the discovery of at least 14 different types of TLR in fish, including TLR9 and TLR21 (25, 26); however, whether these two TLRs are functional has not been investigated previously. In the present study, we comparatively investigated the expression, structural relationship, CpG-ODN interaction, regulation by UNC93B1, and immunologic functions of zebrafish TLR9 (zebTLR9) and TLR21 (zebTLR21) to explore the molecular basis of the immunostimulatory activities of CpG-ODN in fish.  相似文献   

20.
Type I interferons (IFNs) are the first frontline of the host innate immune response against invading pathogens. Herein, we characterized an unknown protein encoded by phospholipase A2 inhibitor and LY6/PLAUR domain-containing (PINLYP) gene that interacted with TBK1 and induced type I IFN in a TBK1- and IRF3-dependent manner. Loss of PINLYP impaired the activation of IRF3 and production of IFN-β induced by DNA virus, RNA virus, and various Toll-like receptor ligands in multiple cell types. Because PINLYP deficiency in mice engendered an early embryonic lethality in mice, we generated a conditional mouse in which PINLYP was depleted in dendritic cells. Mice lacking PINLYP in dendritic cells were defective in type I IFN induction and more susceptible to lethal virus infection. Thus, PINLYP is a positive regulator of type I IFN innate immunity and important for effective host defense against viral infection.

Interferon (IFN)-mediated antiviral responses serve as the first line of the host innate immune defense against viral infection. IFNs are divided into three families based on sequence homology: type I, type II, and type III (1, 2). The type I IFN family encodes 13 subtypes of IFN-α in humans (14 in mice), a single IFN-β subtype, and several poorly defined subtypes (3, 4). Type I IFNs were originally identified based on their ability to interfere with viral replication, restrain virus dissemination, and activate adaptive immune responses (57). They can be induced in most cell types by microbial pathogen-associated and damage-associated molecular patterns recognized by pattern recognition receptors (PRRs) (3). By inducing the expression of IFN-stimulated genes (ISGs), type I IFNs elicit antiviral innate immunity and mediate adaptive immune responses (8, 9).The induction of antiviral type I IFN response is elicited in response to the stimulation of PRRs that detect pathogen-associated molecular patterns, such as viral nucleic acids, viral replicative intermediates, and surface glycoproteins (10, 11). There are four major subfamilies of PRRs: the Toll-like receptors (TLRs), nucleotide-binding oligomerization domain/leucine-rich repeat-containing receptors, RIG-1-like receptors (RLRs), and the C-type lectin receptors, which are located at the cell surface, in the cytosol, or endosomal compartments (1114). Among the TLR family members, TLR3, TLR7, TLR8, and TLR9 are involved in the recognition of viral nucleotides. Viral DNA enriched in CpG-DNA motifs is recognized by TLR9, single-stranded RNA is recognized by TLR7 and TLR8, and double-stranded RNA and its synthetic analog polyinosinic-polycytidylic acid (poly I:C) are recognized by TLR3 (15, 16). Some viral envelope proteins can be recognized by TLR4 or TLR2 (16, 17).Following viral infection, cytosolic DNA can be sensed by cyclic guanosine monophosphate (GMP)–adenosine monophosphate (AMP) synthase (cGAS) that induces the production of cyclic GMP-AMP (cGAMP) (18, 19). cGAMP functions as a second messenger that binds and activates the endoplasmic reticulum (ER) adaptor STING (1922). Translocation of activated STING from the ER to the Golgi apparatus leads to the activation of kinase TBK1, which subsequently phosphorylates IRF3 and triggers the production of type I IFN (2224). Cytosolic RNA can be recognized by the RLRs like RIG-1 and MDA5, which signal via mitochondrial antiviral signaling protein (MAVS; also known as CARDIF, IPS1, and VISA) and subsequently activate TBK1 and IRF3–IRF7, leading to the induction of type I IFNs and other antiviral genes (2527).The lymphocyte antigen-6 (Ly6)/urokinase-type plasminogen activator receptor (uPAR) superfamily is characterized by the LU domain and a domain containing 10 cysteines that form distinct disulfide bridges, which create the three-fingered structural motif. The Ly6/uPAR family members regulate a wide range of functions in various cell types (28). Here, we uncovered the previously uncharacterized role of the Ly6/uPAR family member PINLYP in the induction of type I IFNs in response to DNA virus, RNA virus, and other TLR ligands. This study further defined the pivotal function of PINLYP in the effective host defense against virus infection.  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