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 46 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
The physiological functions of hydrogen sulfide (H2S) include vasorelaxation, stimulation of cellular bioenergetics, and promotion of angiogenesis. Analysis of human colon cancer biopsies and patient-matched normal margin mucosa revealed the selective up-regulation of the H2S-producing enzyme cystathionine-β-synthase (CBS) in colon cancer, resulting in an increased rate of H2S production. Similarly, colon cancer-derived epithelial cell lines (HCT116, HT-29, LoVo) exhibited selective CBS up-regulation and increased H2S production, compared with the nonmalignant colonic mucosa cells, NCM356. CBS localized to the cytosol, as well as the mitochondrial outer membrane. ShRNA-mediated silencing of CBS or its pharmacological inhibition with aminooxyacetic acid reduced HCT116 cell proliferation, migration, and invasion; reduced endothelial cell migration in tumor/endothelial cell cocultures; and suppressed mitochondrial function (oxygen consumption, ATP turnover, and respiratory reserve capacity), as well as glycolysis. Treatment of nude mice with aminooxyacetic acid attenuated the growth of patient-derived colon cancer xenografts and reduced tumor blood flow. Similarly, CBS silencing of the tumor cells decreased xenograft growth and suppressed neovessel density, suggesting a role for endogenous H2S in tumor angiogenesis. In contrast to CBS, silencing of cystathionine-γ-lyase (the expression of which was unchanged in colon cancer) did not affect tumor growth or bioenergetics. In conclusion, H2S produced from CBS serves to (i) maintain colon cancer cellular bioenergetics, thereby supporting tumor growth and proliferation, and (ii) promote angiogenesis and vasorelaxation, consequently providing the tumor with blood and nutritients. The current findings identify CBS-derived H2S as a tumor growth factor and anticancer drug target.The endogenous gasotransmitter hydrogen sulfide (H2S) is a stimulator of vasorelaxation (13), angiogenesis (35), and cellular bioenergetics (6, 7). H2S is generated from l-cysteine by two pyridoxal-5′-phospate–dependent enzymes, cystathionine-β-synthase (CBS) and cystathionine-γ-lyase (CSE), and by the combined action of cysteine aminotransferase and 3-mercaptopyruvate sulfurtransferase (3-MST) (810). H2S exerts its cellular actions via multiple mechanisms (115), including activation of potassium channels (13), stimulation of kinase pathways (4, 11, 12), and inhibition of phosphodiesterases (3, 15).Both ATP generation and angiogenesis are vital factors for the growth and proliferation of tumors (1619). Using human colon cancer tissues and cancer-derived cell lines, we have now conducted a series of in vitro and in vivo studies to explore whether endogenous, tumor cell-derived H2S plays a role as a tumor-derived survival factor. The results show that CBS is selectively overexpressed in colon cancer, and that H2S produced by it serves to maintain the tumor''s cellular bioenergetics and to promote tumor angiogenesis.  相似文献   

6.
7.
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.
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   

10.
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   

11.
We compare the antigenicity and conformation of soluble, cleaved vs. uncleaved envelope glycoprotein (Env gp)140 trimers from the subtype A HIV type 1 (HIV-1) strain BG505. The impact of gp120–gp41 cleavage on trimer structure, in the presence or absence of trimer-stabilizing modifications (i.e., a gp120–gp41 disulfide bond and an I559P gp41 change, together designated SOSIP), was assessed. Without SOSIP changes, cleaved trimers disintegrate into their gp120 and gp41-ectodomain (gp41ECTO) components; when only the disulfide bond is present, they dissociate into gp140 monomers. Uncleaved gp140s remain trimeric whether SOSIP substitutions are present or not. However, negative-stain electron microscopy reveals that only cleaved trimers form homogeneous structures resembling native Env spikes on virus particles. In contrast, uncleaved trimers are highly heterogeneous, adopting a variety of irregular shapes, many of which appear to be gp120 subunits dangling from a central core that is presumably a trimeric form of gp41ECTO. Antigenicity studies with neutralizing and nonneutralizing antibodies are consistent with the EM images; cleaved, SOSIP-stabilized trimers express quaternary structure-dependent epitopes, whereas uncleaved trimers expose nonneutralizing gp120 and gp41ECTO epitopes that are occluded on cleaved trimers. These findings have adverse implications for using soluble, uncleaved trimers for structural studies, and the rationale for testing uncleaved trimers as vaccine candidates also needs to be reevaluated.Trimeric envelope glycoprotein (Env gp) spikes on the HIV type 1 (HIV-1) surface mediate entry of the viral genome into the target cell (1, 2). When spikes interact with their cell-surface receptors, a series of conformational changes within the Env culminates in virus–cell membrane fusion. Neutralizing antibodies (NAbs) against various Env epitopes antagonize these events (2, 3). Hence, Env glycoproteins are a focus of vaccine design programs intended to induce NAbs and thereby prevent HIV-1 transmission (3, 4). Env trimers are composed of three gp120 surface glycoprotein subunits and three gp41 transmembrane glycoproteins, the six subunits all associated via noncovalent interactions (5, 6). A critical event in trimer assembly is proteolytic cleavage of the gp160 precursor into its gp120 and gp41 components, a process essential for HIV-1 entry not least because it liberates the fusion peptide (FP) at the gp41 N terminus (5, 6).Trimer-based vaccine strategies involve expressing soluble, recombinant versions of the virion-associated (i.e., native) spikes. To facilitate production and purification, the membrane-spanning and cytoplasmic domains that anchor spikes to the virion, but that are not NAb targets, are eliminated (712). However, the resulting proteins, known as gp140s, are highly unstable and disintegrate into their gp120 and gp41-ectodomain (gp41ECTO) components, making them useless as immunogens. Two fundamentally different protein-engineering strategies have been used to create gp140s that can be produced and purified without falling apart (3, 4, 717). The most common method involves eliminating the cleavage site between gp120 and gp41ECTO, creating uncleaved gp140s (gp140UNC) where the two subunits remain covalently linked (712). Additional trimerization motifs are often added to the gp41ECTO C terminus (1012). Our alternative approach is based on the premise that cleavage is a fundamental feature of Env structure and involves stabilizing fully cleaved gp140s. The critical changes are an appropriately positioned disulfide bond (referred to as “SOS”) to link gp120 to gp41ECTO covalently, and an Ile/Pro (IP) substitution at residue 559 to strengthen inter-gp41ECTO interactions (1317). The resulting cleaved trimers are designated SOSIP gp140s (14). Additional modifications have improved their stability, homogeneity, and antigenicity (1517). Our current design, based on the BG505 subtype A env gene, yields SOSIP.664 trimers that mimic native, virion-associated Env spikes antigenically and when viewed by negative-stain electron microscopy (EM) (1719).Here we show that cleavage is essential for producing stable, soluble gp140 trimers that resemble native Env spikes. EM studies reveal that purified, trimeric gp140UNC proteins are heterogeneous and that the irregularly shaped images rarely resemble a native spike; we refer to them as “aberrant configurations” (ACs). In contrast, cleaved SOSIP gp140 trimers are homogeneous and mimic native spikes; we designate them native-like (NL) trimers. The antigenic properties of the cleaved (NL) and uncleaved (AC) trimers, assessed by surface plasmon resonance (SPR) and enzyme-linked immunoabsorbance assays (ELISA), are consistent with the EM images. Nonneutralizing gp120 and gp41ECTO epitopes are exposed on gp140UNC trimers but occluded on cleaved ones, whereas quaternary structure-dependent epitopes indicative of proper folding are present only on cleaved trimers. Our findings have substantial implications, because uncleaved trimers are being studied structurally and developed as vaccine candidates (3, 9, 10, 12, 20).  相似文献   

12.
cAMP is an evolutionary conserved, prototypic second messenger regulating numerous cellular functions. In mammals, cAMP is synthesized by one of 10 homologous adenylyl cyclases (ACs): nine transmembrane enzymes and one soluble AC (sAC). Among these, only sAC is directly activated by bicarbonate (HCO3); it thereby serves as a cellular sensor for HCO3, carbon dioxide (CO2), and pH in physiological functions, such as sperm activation, aqueous humor formation, and metabolic regulation. Here, we describe crystal structures of human sAC catalytic domains in the apo state and in complex with substrate analog, products, and regulators. The activator HCO3 binds adjacent to Arg176, which acts as a switch that enables formation of the catalytic cation sites. An anionic inhibitor, 4,4′-diisothiocyanatostilbene-2,2′-disulfonic acid, inhibits sAC through binding to the active site entrance, which blocks HCO3 activation through steric hindrance and trapping of the Arg176 side chain. Finally, product complexes reveal small, local rearrangements that facilitate catalysis. Our results provide a molecular mechanism for sAC catalysis and cellular HCO3 sensing and a basis for targeting this system with drugs.The ubiquitous second messenger cAMP regulates diverse physiological processes, from fungal virulence to mammalian brain function (1, 2). In mammals, cAMP can be generated by any of 10 differently expressed and regulated adenylyl cyclases (ACs): nine transmembrane enzymes (tmACs) and one soluble AC (sAC) (3). TmACs reside in the cell membrane, where they mediate cellular responses to hormones acting through G protein-coupled receptors (4). In contrast, sAC functions in various intracellular locations, providing cell-specific spatial and temporal patterns of cAMP (57) in response to intracellular signals, including calcium, ATP, and bicarbonate (HCO3) (3, 810). HCO3 regulation of sAC enzymes is a direct effect on their catalytic domains and is conserved across bacterial, fungal, and animal kingdoms (1, 1113). Via modulation of sAC, and sAC-like cyclase activities, HCO3 serves as an evolutionarily conserved signaling molecule mediating cellular responses to HCO3, CO2, and pH (3, 14). In mammals, sAC acts as a CO2/HCO3/pH sensor in processes such as sperm activation (15), acid-base homeostasis (16), and various metabolic responses (10, 17, 18). sAC has also been implicated in skin and prostate cancer and as a target for male contraceptives (1921).All mammalian ACs are class III nucleotidyl cyclases sharing homologous catalytic domains. Their catalytic cores are formed through symmetrical or pseudosymmetrical association of two identical or highly similar catalytic domains, C1 and C2 (2224); in mammalian ACs, both domains reside on a single polypeptide chain. Such C1C2 pseudoheterodimers form two pseudosymmetrical sites at the dimer interface: the active site and a degenerated, inactive pocket (3, 23). A conserved Lys and an Asp/Thr in the active site recognize the base of the substrate ATP, and two conserved Asp residues bind two divalent cations, normally Mg2+ (23). The ions, called ion A and ion B, coordinate the substrate phosphates and support the intramolecular 3′-hydroxyl (3′-OH) attack at the α-phosphorous to form cAMP and pyrophosphate (PPi) (3). In tmACs, the degenerate site binds forskolin (24), a plant diterpene that activates tmACs but has no effect on sAC (25). The forskolin activation mechanism and the existence of physiological ligands for this site in tmACs or in sAC remain unclear.There are two sAC isoforms known to be generated by alternative splicing (26). Full-length sAC comprises N-terminal catalytic domains along with ∼1,100 residues with a little understood function except for an autoinhibitory motif and a heme-binding domain (3, 27, 28). Exclusion of exon 12 (26) generates a truncated isoform, sACt (residues 1–490), which comprises just the two sAC catalytic domains (sAC-cat) (25). sACt is widely expressed, and it is the isoform most extensively biochemically characterized (3, 8, 11). It is directly activated by Ca2+ and HCO3; Ca2+ supports substrate binding, and HCO3 increases turnover and relieves substrate inhibition (8), and this regulation is conserved in sAC-like enzymes from Cyanobacteria to humans (3, 13, 29). In a homodimeric, HCO3-regulated sAC homolog from Spirulina platensis, adenylyl cyclase C (CyaC), HCO3 appeared to facilitate an active site closure required for catalysis (13), but the HCO3 binding site and its mechanism of activation remained unknown.Here, we present crystal structures of the human sAC-cat in apo form and in complex with substrate, products, bicarbonate, and a pharmacological inhibitor. The structures reveal insights into binding sites and mechanisms for sAC catalysis and for its regulation by physiological and pharmacological small molecules.  相似文献   

13.
Mutations that lead to Huntington’s disease (HD) result in increased transmission at glutamatergic corticostriatal synapses at early presymptomatic stages that have been postulated to set the stage for pathological changes and symptoms that are observed at later ages. Based on this, pharmacological interventions that reverse excessive corticostriatal transmission may provide a novel approach for reducing early physiological changes and motor symptoms observed in HD. We report that activation of the M4 subtype of muscarinic acetylcholine receptor reduces transmission at corticostriatal synapses and that this effect is dramatically enhanced in presymptomatic YAC128 HD and BACHD relative to wild-type mice. Furthermore, chronic administration of a novel highly selective M4 positive allosteric modulator (PAM) beginning at presymptomatic ages improves motor and synaptic deficits in 5-mo-old YAC128 mice. These data raise the exciting possibility that selective M4 PAMs could provide a therapeutic strategy for the treatment of HD.Huntington’s disease (HD) is a rare and fatal neurodegenerative disease caused by an expansion of a CAG triplet repeat in Htt, the gene that encodes for the protein huntingtin (1, 2). HD is characterized by a prediagnostic phase that includes subtle changes in personality, cognition, and motor function, followed by a more severe symptomatic stage initially characterized by hyperkinesia (chorea), motor incoordination, deterioration of cognitive abilities, and psychiatric symptoms. At later stages of disease progression, patients experience dystonia, rigidity, and bradykinesia, and ultimately death (37). The cortex and striatum are the most severely affected brain regions in HD and, interestingly, an increasing number of reports suggest that alterations in cortical and striatal physiology are present in prediagnostic individuals and in young HD mice (616).Striatal spiny projection neurons (SPNs) receive large glutamatergic inputs from the cortex and thalamus, as well as dopaminergic innervation from the substantia nigra. In the healthy striatum, the interplay of these neurotransmitters coordinates the activity of SPNs and striatal interneurons, regulating motor planning and execution as well as cognition and motivation (17, 18). Htt mutations lead to an early increase in striatal glutamatergic transmission, which begins during the asymptomatic phase of HD (1214) and could contribute to synaptic changes observed in later stages of HD (19, 20). Based on this, pharmacological agents that reduce excitatory transmission in the striatum could reduce or prevent the progression of alterations in striatal synaptic function and behavior observed in symptomatic stages of HD.Muscarinic acetylcholine receptors (mAChRs), particularly M4, can inhibit transmission at corticostriatal synapses (2125). Therefore, it is possible that selective activation of specific mAChR subtypes could normalize excessive corticostriatal transmission in HD. Interestingly, previous studies also suggest that HD is associated with alterations of striatal cholinergic markers, including mAChRs (2629). We now provide exciting new evidence that M4-mediated control of corticostriatal transmission is increased in young asymptomatic HD mice and that M4 positive allosteric modulators (PAMs) may represent a new treatment strategy for normalizing early changes in corticostriatal transmission and reducing the progression of HD.  相似文献   

14.
15.
The increase of extracellular heme is a hallmark of hemolysis or extensive cell damage. Heme has prooxidant, cytotoxic, and inflammatory effects, playing a central role in the pathogenesis of malaria, sepsis, and sickle cell disease. However, the mechanisms by which heme is sensed by innate immune cells contributing to these diseases are not fully characterized. We found that heme, but not porphyrins without iron, activated LPS-primed macrophages promoting the processing of IL-1β dependent on nucleotide-binding domain and leucine rich repeat containing family, pyrin domain containing 3 (NLRP3). The activation of NLRP3 by heme required spleen tyrosine kinase, NADPH oxidase-2, mitochondrial reactive oxygen species, and K+ efflux, whereas it was independent of heme internalization, lysosomal damage, ATP release, the purinergic receptor P2X7, and cell death. Importantly, our results indicated the participation of macrophages, NLRP3 inflammasome components, and IL-1R in the lethality caused by sterile hemolysis. Thus, understanding the molecular pathways affected by heme in innate immune cells might prove useful to identify new therapeutic targets for diseases that have heme release.Hemolysis, hemorrhage, and rhabdomyolysis cause the release of large amounts of hemoproteins to the extracellular space, which, once oxidized, release the heme moiety, a potentially harmful molecule due to its prooxidant, cytotoxic, and inflammatory effects (1, 2). Scavenging proteins such as haptoglobin and hemopexin bind hemoglobin and heme, respectively, promoting their clearance from the circulation and delivery to cells involved with heme catabolism. Heme oxygenase cleaves heme and generates equimolar amounts of biliverdin, carbon monoxide (CO) and iron (2). Studies using mice deficient for haptoglobin (Hp), hemopexin (Hx), and heme oxygenase 1 (HO-1) demonstrate the importance of these proteins in controlling the deleterious effects of heme. Both Hp−/− and Hx−/− mice have increased renal damage after acute hemolysis induced by phenyhydrazine (Phz) compared with wild-type mice (3, 4). Mice lacking both proteins present splenomegaly and liver inflammation composed of several foci with leukocyte infiltration after intravascular hemolysis (5). Hx protect mice against heme-induced endothelial damage improving liver and cardiovascular function (68). Lack of heme oxygenase 1 (Hmox1−/−) causes iron overload, increased cell death, and tissue inflammation under basal conditions and upon inflammatory stimuli (915). This salutary effect of HO-1 has been attributed to its effect of reducing heme amounts as well as generating the cytoprotective molecules, biliverdin and CO.Heme induces neutrophil migration in vivo and in vitro (16, 17), inhibits neutrophil apoptosis (18), triggers cytokine and lipid mediator production by macrophages (19, 20), and increases the expression of adhesion molecules and tissue factor on endothelial cells (2123). Heme cooperates with TNF, causing hepatocyte apoptosis in a mechanism dependent on reactive oxygen species (ROS) generation (12). Whereas heme-induced TNF production depends on functional toll-like receptor 4 (TLR4), ROS generation in response to heme is TLR4 independent (19). We recently observed that heme triggers receptor-interacting protein (RIP)1/3-dependent macrophage-programmed necrosis through the induction of TNF and ROS (15). The highly unstable nature of iron is considered critical for the ability of heme to generate ROS and to cause inflammation. ROS generated by heme has been mainly attributed to the Fenton reaction. However, recent studies suggest that heme can generate ROS through multiple sources, including NADPH oxidase and mitochondria (22, 2427).Heme causes inflammation in sterile and infectious conditions, contributing to the pathogenesis of hemolytic diseases, subarachnoid hemorrhage, malaria, and sepsis (11, 13, 24, 28), but the mechanisms by which heme operates in different conditions are not completely understood. Blocking the prooxidant effects of heme protects cells from death and prevents tissue damage and lethality in models of malaria and sepsis (12, 13, 15). Importantly, two recent studies demonstrated the pathogenic role of heme-induced TLR4 activation in a mouse model of sickle cell disease (29, 30). These results highlight the great potential of understanding the molecular mechanisms of heme-induced inflammation and cell death as a way to identify new therapeutic targets.Hemolysis and heme synergize with microbial molecules for the induction of inflammatory cytokine production and inflammation in a mechanism dependent on ROS and Syk (24). Processing of pro–IL-1β is dependent on caspase-1 activity, requiring assembly of the inflammasome, a cytosolic multiprotein complex composed of a NOD-like receptor, the adaptor protein apoptosis-associated speck-like protein containing a CARD (ASC), and caspase-1 (3133). The most extensively studied inflammasome is the nucleotide-binding domain and leucine rich repeat containing family, pyrin domain containing 3 (NLRP3). NLRP3 and pro–IL-1β expression are increased in innate immune cells primed with NF-κB inducers such as TLR agonists and TNF (34, 35). NLRP3 inflammasome is activated by several structurally nonrelated stimuli, such as endogenous and microbial molecules, pore-forming toxins, and particulate matter (34, 35). The activation of NLRP3 involves K+ efflux, increase of ROS and Syk phosphorylation. Importantly, critical roles of NLRP3 have been demonstrated in a vast number of diseases (34, 36). We hypothesize that heme causes the activation of the inflammasome and secretion of IL-1β. Here we found that heme triggered the processing and secretion of IL-1β dependently on NLRP3 inflammasome in vitro and in vivo. The activation of NLRP3 by heme was dependent on Syk, ROS, and K+ efflux, but independent of lysosomal leakage, ATP release, or cell death. Finally, our results indicated the critical role of macrophages, the NLRP3 inflammasome, and IL-1R to the lethality caused by sterile hemolysis.  相似文献   

16.
Fundamental relationships between the thermodynamics and kinetics of protein folding were investigated using chain models of natural proteins with diverse folding rates by extensive comparisons between the distribution of conformations in thermodynamic equilibrium and the distribution of conformations sampled along folding trajectories. Consistent with theory and single-molecule experiment, duration of the folding transition paths exhibits only a weak correlation with overall folding time. Conformational distributions of folding trajectories near the overall thermodynamic folding/unfolding barrier show significant deviations from preequilibrium. These deviations, the distribution of transition path times, and the variation of mean transition path time for different proteins can all be rationalized by a diffusive process that we modeled using simple Monte Carlo algorithms with an effective coordinate-independent diffusion coefficient. Conformations in the initial stages of transition paths tend to form more nonlocal contacts than typical conformations with the same number of native contacts. This statistical bias, which is indicative of preferred folding pathways, should be amenable to future single-molecule measurements. We found that the preexponential factor defined in the transition state theory of folding varies from protein to protein and that this variation can be rationalized by our Monte Carlo diffusion model. Thus, protein folding physics is different in certain fundamental respects from the physics envisioned by a simple transition-state picture. Nonetheless, transition state theory can be a useful approximate predictor of cooperative folding speed, because the height of the overall folding barrier is apparently a proxy for related rate-determining physical properties.Protein folding is an intriguing phenomenon at the interface of physics and biology. In the early days of folding kinetics studies, folding was formulated almost exclusively in terms of mass-action rate equations connecting the folded, unfolded, and possibly, one or a few intermediate states (1, 2). With the advent of site-directed mutagenesis, the concept of free energy barriers from transition state theory (TST) (3) was introduced to interpret mutational data (4), and subsequently, it was adopted for the Φ-value analysis (5). Since the 1990s, the availability of more detailed experimental data (6), in conjunction with computational development of coarse-grained chain models, has led to an energy landscape picture of folding (715). This perspective emphasizes the diversity of microscopic folding trajectories, and it conceptualizes folding as a diffusive process (1625) akin to the theory of Kramers (26).For two-state-like folding, the transition path (TP), i.e., the sequence of kinetic events that leads directly from the unfolded state to the folded state (27, 28), constitutes only a tiny fraction of a folding trajectory that spends most of the time diffusing, seemingly unproductively, in the vicinity of the free energy minimum of the unfolded state. The development of ultrafast laser spectroscopy (29, 30) and single-molecule (27, 28, 31) techniques have made it possible to establish upper bounds on the transition path time (tTP) ranging from <200 and <10 μs by earlier (27) and more recent (28), respectively, direct single-molecule FRET to <2 μs (30) by bulk relaxation measurements. Consistent with these observations, recent extensive atomic simulations have also provided estimated tTP values of the order of ∼1 μs (32, 33). These advances offer exciting prospects of characterizing the productive events along folding TPs.It is timely, therefore, to further the theoretical investigation of TP-related questions (19). To this end, we used coarse-grained Cα models (14) to perform extensive simulations of the folding trajectories of small proteins with 56- to 86-aa residues. These tractable models are useful, because despite significant progress, current atomic models cannot provide the same degree of sampling coverage for proteins of comparable sizes (32, 33). In addition to structural insights, this study provides previously unexplored vantage points to compare the diffusion and TST pictures of folding. Deviations of folding behaviors from TST predictions are not unexpected, because TST is mostly applicable to simple gas reactions; however, the nature and extent of the deviations have not been much explored. Our explicit-chain simulation data conform well to the diffusion picture but not as well to TST. In particular, the preexponential factors of the simulated folding rates exhibit a small but appreciable variation that depends on native topology. These findings and others reported below underscore the importance of single-molecule measurements (13, 27, 28, 31, 34, 35) in assessing the merits of proposed scenarios and organizing principles of folding (725, 36, 37).  相似文献   

17.
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   

18.
The inositol 1,4,5-trisphosphate receptor (IP3R) in the endoplasmic reticulum mediates calcium signaling that impinges on intracellular processes. IP3Rs are allosteric proteins comprising four subunits that form an ion channel activated by binding of IP3 at a distance. Defective allostery in IP3R is considered crucial to cellular dysfunction, but the specific mechanism remains unknown. Here we demonstrate that a pleiotropic enzyme transglutaminase type 2 targets the allosteric coupling domain of IP3R type 1 (IP3R1) and negatively regulates IP3R1-mediated calcium signaling and autophagy by locking the subunit configurations. The control point of this regulation is the covalent posttranslational modification of the Gln2746 residue that transglutaminase type 2 tethers to the adjacent subunit. Modification of Gln2746 and IP3R1 function was observed in Huntington disease models, suggesting a pathological role of this modification in the neurodegenerative disease. Our study reveals that cellular signaling is regulated by a new mode of posttranslational modification that chronically and enzymatically blocks allosteric changes in the ligand-gated channels that relate to disease states.Ligand-gated ion channels function by allostery that is the regulation at a distance; the allosteric coupling of ligand binding with channel gating requires reversible changes in subunit configurations and conformations (1). Inositol 1,4,5-trisphosphate receptors (IP3Rs) are ligand-gated ion channels that release calcium ions (Ca2+) from the endoplasmic reticulum (ER) (2, 3). IP3Rs are allosteric proteins comprising four subunits that assemble a calcium channel with fourfold symmetry about an axis perpendicular to the ER membrane. The subunits of three IP3R isoforms (IP3R1, IP3R2, and IP3R3) are structurally divided into three domains: the IP3-binding domain (IBD), the regulatory domain, and the channel domain (36). Fitting of the IBD X-ray structures (7, 8) to a cryo-EM map (9) indicates that the IBD activates a remote Ca2+ channel by allostery (8); however, the current X-ray structure only spans 5% of each tetramer, such that the mechanism underlying allosteric coupling of the IBD to channel gating remains unknown.The IP3R in the ER mediates intracellular calcium signaling that impinges on homeostatic control in various subsequent intracellular processes. Deletion of the genes encoding the type 1 IP3R (IP3R1) leads to perturbations in long-term potentiation/depression (3, 10, 11) and spinogenesis (12), and the human genetic disease spinocerebellar ataxia 15 is caused by haploinsufficiency of the IP3R1 gene (1315). Dysregulation of IP3R1 is also implicated in neurodegenerative diseases including Huntington disease (HD) (1618) and Alzheimer’s disease (AD) (1922). IP3Rs also control fundamental cellular processes—for example, mitochondrial energy production (23, 24), autophagy regulation (2427), ER stress (28), hepatic gluconeogenesis (29), pancreatic exocytosis (30), and macrophage inflammasomes (31). On the other hand, excessive IP3R function promotes cell death processes including apoptosis by activating mitochondrial or calpain pathways (2, 17). Considering these versatile roles of IP3Rs, appropriate IP3R structure and function are essential for living systems, and aberrant regulation of IP3R closely relates to various diseases.Several factors such as cytosolic molecules, interacting proteins, and posttranslational modifications control the IP3-induced Ca2+ release (IICR) through allosteric sites in IP3Rs. Cytosolic Ca2+ concentrations strictly control IICR in a biphasic manner with activation at low concentrations and inhibition at higher concentrations. The critical Ca2+ sensor for activation is conserved among the three isoforms of IP3 and ryanodine receptors, and this sensor is located in the regulatory domain outside the IBD and the channel domain (32). A putative ATP regulatory region is deleted in opisthotonos mice, and IICR is also regulated by this mutation in the regulatory domain (33). Various interacting proteins, such as cytochrome c, Bcl-2-family proteins, ataxin-3, huntingtin (Htt) protein, Htt-associated protein 1A (HAP1A), and G-protein–coupled receptor kinase-interacting protein 1 (GIT1), target allosteric sites in the carboxyl-terminal tail (35). The regulatory domain and the carboxyl-terminal tail also undergo phosphorylation by the protein kinases A/G and B/Akt and contain the apoptotic cleavage sites for the protease caspase-3 (4, 5). These factors allosterically regulate IP3R structure and function to control cellular fates; therefore, understanding the allosteric coupling of the IBD to channel gating will elucidate the regulatory mechanism of these factors.Transglutaminase (TG) catalyses protein cross-linking between a glutamine (Gln) residue and a lysine (Lys) residue via an Nε-(γ-glutamyl)lysine isopeptide bond (34, 35). TG type 2 (TG2) is a Ca2+-dependent enzyme with widespread distribution and is highly inducible by various stimulations such as oxidative stress, cytokines, growth factors, and retinoic acid (RA) (34, 35). TG2 is considered a significant disease-modifying factor in neurodegenerative diseases including HD, AD, and Parkinson’s diseases (PD) (34, 3645) because TG2 might enzymatically stabilize aberrant aggregates of proteins implicated in these diseases—that is, mutant Htt, β-amyloid, and α-synuclein; however, the causal role of TG2 in Ca2+ signaling in brain pathogenesis has been unclear. Ablation of TG2 in HD mouse models is associated with increased lifespan and improved motor function (46, 47). However, TG2 knockout mice do not show impaired Htt aggregation, suggesting that TG2 may play a causal role in these disorders rather than TG2-dependent cross-links in aberrant protein aggregates (47, 48).In this study, we discovered a new mode of chronic and irreversible allosteric regulation in IP3R1 in which covalent modification of the receptor at Gln2746 is catalyzed by TG2. We demonstrate that up-regulation of TG2 modifies IP3R1 structure and function in HD models and propose an etiologic role of this modification in the reduction of neuronal signaling and subsequent processes during the prodromal state of the neurodegenerative disease.  相似文献   

19.
In vitro evidence suggests that plasmacytoid dendritic cells (pDCs) are intimately involved in the pathogenesis of lupus. However, it remains to be determined whether these cells are required in vivo for disease development, and whether their contribution is restricted to hyperproduction of type I IFNs. To address these issues, we created lupus-predisposed mice lacking the IFN regulatory factor 8 (IRF8) or carrying a mutation that impairs the peptide/histidine transporter solute carrier family 15, member 4 (SLC15A4). IRF8-deficient NZB mice, lacking pDCs, showed almost complete absence of anti-nuclear, anti-chromatin, and anti-erythrocyte autoantibodies, along with reduced kidney disease. These effects were observed despite normal B-cell responses to Toll-like receptor (TLR) 7 and TLR9 stimuli and intact humoral responses to conventional T-dependent and -independent antigens. Moreover, Slc15a4 mutant C57BL/6-Faslpr mice, in which pDCs are present but unable to produce type I IFNs in response to endosomal TLR ligands, also showed an absence of autoantibodies, reduced lymphadenopathy and splenomegaly, and extended survival. Taken together, our results demonstrate that pDCs and the production of type I IFNs by these cells are critical contributors to the pathogenesis of lupus-like autoimmunity in these models. Thus, IRF8 and SLC15A4 may provide important targets for therapeutic intervention in human lupus.Extensive evidence suggests that type I IFNs are major pathogenic effectors in lupus-associated systemic autoimmunity. A well-documented pattern of expression of type I IFN-inducible genes occurs in peripheral blood mononuclear cells of patients with systemic lupus erythematosus (SLE) (13), and reduced disease is observed in some lupus-predisposed mice that either lack the common receptor (IFNAR) for these cytokines (4, 5) or have been treated with IFNAR-blocking antibody (6). Consequently, attention has focused on defining the cell subsets and signaling processes involved in type I IFN production, the mechanisms by which these mediators accelerate disease, and approaches to interfere with these pathogenic events.Early in vitro studies showed that type I IFN production can be induced in normal blood leukocytes by SLE autoantibodies complexed with nucleic acid-containing apoptotic/necrotic cell material, and further work demonstrated that this activity is sensitive to RNase and DNase digestion (7, 8). These results were integrated in a more comprehensive scheme following the demonstration that type I IFN induction by these complexes is mediated by the engagement of endosomal Toll-like receptors (TLRs) (911). Similarly, antigenic cargo containing nucleic acids was found to promote B-cell proliferation in a TLR9- or TLR7-dependent manner, with this effect enhanced by type I IFN signaling (9, 12, 13). The contribution of nucleic acid-sensing TLRs to systemic autoimmunity was further corroborated by studies in lupus-predisposed mice lacking or overexpressing TLR7 and/or TLR9 (14-20), and in Unc93b1 (3d) mutant mice in which signaling by endosomal TLRs is extinguished (21).The cell population involved in type I IFN production in response to lupus-related immune complexes corresponds to natural IFN-producing cells (22, 23). These cells, known as plasmacytoid DCs (pDCs), are the most potent producers of type I IFNs, a functional characteristic attributed to constitutive expression of TLR7, TLR9, and IRF7 and likely signaling from a unique intracellular compartment (2427). The involvement of pDCs in lupus is further suggested by the reduced frequency of these cells in patient blood together with increases in afflicted organs, presumably caused by the attraction of activated pDCs to inflammatory sites (10). Similar increases have been noted in inflammatory tissues of patients with Sjögren''s syndrome (28), rheumatoid arthritis (29, 30), dermatomyositis (31), and psoriasis (32).Collectively, these results suggest that pDCs, acting through type I IFN hyperproduction, are major pathogenic contributors to lupus. Whether the participation of these cells is obligatory remains to be documented in vivo, however. Here, using congenic lupus-predisposed mice lacking pDCs (as well as other DC subsets) owing to IRF8 deficiency, or exhibiting pDC-specific defects in endosomal TLR signaling and type I IFN production owing to Slc15a4 (feeble) mutation, we provide strong evidence that pDCs are indeed required for disease development, and this effect appears to be mediated by hyperproduction of inflammatory cytokines, most likely type I IFNs.  相似文献   

20.
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