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 406 毫秒
1.
T lymphocytes are characterized by the use of structurally diverse TCR. The discovery of subsets of canonical T cells that have structurally homogeneous TCR presents an enigma: What antigens do these T cells recognize, and how does their antigen specificity relate to their functions? One subset of canonical T cells is restricted by CD1d, a non-classical antigen presenting molecule that presents lipids and glycolipids. Canonical CD1d-restricted T cells have semi-invariant TCR consisting of an invariantly rearranged TCR alpha chain, paired with diversely rearranged TCR beta chains. Most respond strongly to the unusual glycolipid alpha-galactosylceramide (alpha-GalCer), and can also respond to cellular antigens presented by CD1d. Mounting evidence indicates that alpha-GalCer responsive T cells are heterogeneous in their reactivities to cellular antigens, suggesting that an individual semi-invariant TCR may be capable of recognizing more than one ligand. Recent crystal structures of CD1b molecules with three different bound lipids indicate that the antigenic features of lipids may be localized over a smaller area than those of peptides, and that the positioning of the polar head group can vary substantially. A model that explains how CD1d-restricted T cells could possess both conserved and heterogeneous antigen specificities, is that different lipid antigens may interact with distinct areas of a TCR due to differences in the positioning of the polar head group. Hence, canonical CD1d-restricted TCR could recognize conserved antigens via the invariant TCR alpha chain, and have diverse antigen specificities that are conferred by their individual TCR beta chains.  相似文献   

2.
Postsynaptic neuroligins are thought to perform essential functions in synapse validation and synaptic transmission by binding to, and dimerizing, presynaptic α‐ and β‐neurexins. To test this hypothesis, we examined the functional effects of neuroligin‐1 mutations that impair only α‐neurexin binding, block both α‐ and β‐neurexin binding, or abolish neuroligin‐1 dimerization. Abolishing α‐neurexin binding abrogated neuroligin‐induced generation of neuronal synapses onto transfected non‐neuronal cells in the so‐called artificial synapse‐formation assay, even though β‐neurexin binding was retained. Thus, in this assay, neuroligin‐1 induces apparent synapse formation by binding to presynaptic α‐neurexins. In transfected neurons, however, neither α‐ nor β‐neurexin binding was essential for the ability of postsynaptic neuroligin‐1 to dramatically increase synapse density, suggesting a neurexin‐independent mechanism of synapse formation. Moreover, neuroligin‐1 dimerization was not required for either the non‐neuronal or the neuronal synapse‐formation assay. Nevertheless, both α‐neurexin binding and neuroligin‐1 dimerization were essential for the increase in apparent synapse size that is induced by neuroligin‐1 in transfected neurons. Thus, neuroligin‐1 performs diverse synaptic functions by mechanisms that include as essential components of α‐neurexin binding and neuroligin dimerization, but extend beyond these activities.  相似文献   

3.
Plants offer fast, flexible and easily scalable alternative platforms for the production of pharmaceutical proteins, but differences between plant and mammalian N‐linked glycans, including the presence of β‐1,2‐xylose and core α‐1,3‐fucose residues in plants, can affect the activity, potency and immunogenicity of plant‐derived proteins. Nicotiana benthamiana is widely used for the transient expression of recombinant proteins so it is desirable to modify the endogenous N‐glycosylation machinery to allow the synthesis of complex N‐glycans lacking β‐1,2‐xylose and core α‐1,3‐fucose. Here, we used multiplex CRISPR/Cas9 genome editing to generate N. benthamiana production lines deficient in plant‐specific α‐1,3‐fucosyltransferase and β‐1,2‐xylosyltransferase activity, reflecting the mutation of six different genes. We confirmed the functional gene knockouts by Sanger sequencing and mass spectrometry‐based N‐glycan analysis of endogenous proteins and the recombinant monoclonal antibody 2G12. Furthermore, we compared the CD64‐binding affinity of 2G12 glycovariants produced in wild‐type N. benthamiana, the newly generated FX‐KO line, and Chinese hamster ovary (CHO) cells, confirming that the glyco‐engineered antibody performed as well as its CHO‐produced counterpart.  相似文献   

4.
Introduction – Bioautographic assays using TLC play an important role in the search for active compounds from plants. A TLC assay has previously been established for the detection of β‐glucosidase inhibitors but not for α‐glucosidase. Nonetheless, α‐glucosidase inhibition is an important target for therapeutic agents against of type 2 diabetes and anti‐viral infections. Objective – To develop a TLC bioautographic method to detect α‐ and β‐glucosidase inhibitors in plant extracts. Methodology – The enzymes α‐ and β‐d ‐glucosidase were dissolved in sodium acetate buffer. After migration of the samples, the TLC plate was sprayed with enzyme solution and incubated at room temperature for 60 min in the case of α‐d ‐glucosidase, and 37°C for 20 min in the case of β‐d ‐glucosidase. For detection of the active enzyme, solutions of 2‐naphthyl‐α‐D‐glucopyranoside or 2‐naphthyl‐β‐D‐glucopyranoside and Fast Blue Salt were mixed at a ratio of 1 : 1 (for α‐d ‐glucosidase) or 1 : 4 (for β‐d ‐glucosidase) and sprayed onto the plate to give a purple background colouration after 2–5 min. Results – Enzyme inhibitors were visualised as white spots on the TLC plates. Conduritol B epoxide inhibited α‐d ‐glucosidase and β‐d ‐glucosidase down to 0.1 µg. Methanol extracts of Tussilago farfara and Urtica dioica after migration on TLC gave enzymatic inhibition when applied in amounts of 100 µg for α‐glucosidase and 50 µg for β‐glucosidase. Conclusion – The screening test was able to detect inhibition of α‐ and β‐glucosidases by pure reference substances and by compounds present in complex matrices, such as plant extracts. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

5.
CD1d function is regulated by microsomal triglyceride transfer protein   总被引:7,自引:0,他引:7  
CD1d is a major histocompatibility complex (MHC) class I-related molecule that functions in glycolipid antigen presentation to distinct subsets of T cells that express natural killer receptors and an invariant T-cell receptor-alpha chain (invariant NKT cells). The acquisition of glycolipid antigens by CD1d occurs, in part, in endosomes through the function of resident lipid transfer proteins, namely saposins. Here we show that microsomal triglyceride transfer protein (MTP), a protein that resides in the endoplasmic reticulum of hepatocytes and intestinal epithelial cells (IECs) and is essential for lipidation of apolipoprotein B, associates with CD1d in hepatocytes. Hepatocytes from animals in which Mttp (the gene encoding MTP) has been conditionally deleted, and IECs in which Mttp gene products have been silenced, are unable to activate invariant NKT cells. Conditional deletion of the Mttp gene in hepatocytes is associated with a redistribution of CD1d expression, and Mttp-deleted mice are resistant to immunopathologies associated with invariant NKT cell-mediated hepatitis and colitis. These studies indicate that the CD1d-regulating function of MTP in the endoplasmic reticulum is complementary to that of the saposins in endosomes in vivo.  相似文献   

6.
HIV‐1 Vpu modulates cellular transmembrane proteins to optimize viral replication and provide immune‐evasion, triggering ubiquitin‐mediated degradation of some targets but also modulating endosomal trafficking to deplete them from the plasma membrane. Interactions between Vpu and the heterotetrameric clathrin adaptor protein (AP) complexes AP‐1 and AP‐2 have been described, yet the molecular basis and functional roles of such interactions are incompletely defined. To investigate the trafficking signals encoded by Vpu, we fused the cytoplasmic domain (CD) of Vpu to the extracellular and transmembrane domains of the CD8 α‐chain. CD8‐VpuCD was rapidly endocytosed in a clathrin‐ and AP‐2‐dependent manner. Multiple determinants within the Vpu CD contributed to endocytic activity, including phosphoserines of the β‐TrCP binding site and a leucine‐based ExxxLV motif. Using recombinant proteins, we confirmed ExxxLV‐dependent binding of the Vpu CD to the α/σ2 subunit hemicomplex of AP‐2 and showed that this is enhanced by serine‐phosphorylation. Remarkably, the Vpu CD also bound directly to the medium (μ) subunits of AP‐2 and AP‐1; this interaction was dependent on serine‐phosphorylation of Vpu and on basic residues in the μ subunits. We propose that the flexibility with which Vpu binds AP complexes broadens the range of cellular targets that it can misdirect to the virus' advantage.   相似文献   

7.
Increased expression of T cell immunoglobulin and mucin domain‐3 (Tim‐3) on invariant natural killer T (iNKT) cells is reported in chronic hepatitis B virus (HBV) infection. However, whether Tim‐3 regulates iNKT cells in chronic HBV condition remains unclear. In this study, our results showed that the expression of Tim‐3 was up‐regulated on hepatic iNKT cells from HBV‐transgenic (Tg) mice or iNKT cells stimulated with α‐galactosylceramide (α‐Galcer). Compared with Tim‐3?iNKT cells, Tim‐3+iNKT cells expressed more IFN‐γ, IL‐4 and CD107a, indicating a strong relationship between Tim‐3 and iNKT cell activation. Constantly, treatment of Tim‐3 blocking antibodies significantly enhanced the production of IFN‐γ, TNF‐α, IL‐4 and CD107a in iNKT cells both in vivo and in vitro. This Tim‐3? mediated suppression of iNKT cells was further confirmed in Tim‐3 knockout (KO) mice. Moreover, Tim‐3 blockade promoted α‐Galcer‐triggered inhibition of HBV replication, displaying as the decreased HBV DNA and HBsAg level in serum, and down‐regulated pgRNA expression in liver tissues. Collectively, our data, for the first time, demonstrated the potential role of Tim‐3 blockade in promoting iNKT cell‐mediated HBV inhibition. Therefore, combination of α‐Galcer with Tim‐3 blockade might be a promising approach in chronic hepatitis B therapy.  相似文献   

8.
Protein spin labeling to yield the nitroxide‐based R1 side chain is a powerful method to measure protein dynamics and structure by electron spin resonance. However, R1 measurements are complicated by the flexibility of the side chain. While analysis approaches for solvent‐exposed α‐helical environment have been developed to partially account for flexibility, similar work in β‐sheets is lacking. The goal of this study is to provide the first essential steps for understanding the conformational preferences of R1 within edge β‐strands using X‐ray crystallography and double electron electron resonance (DEER) distance measurements. Crystal structures yielded seven rotamers for a non‐hydrogen‐bonded site and three rotamers for a hydrogen‐bonded site. The observed rotamers indicate contextual differences in R1 conformational preferences compared to other solvent‐exposed environments. For the DEER measurements, each strand site was paired with the same α‐helical site elsewhere on the protein. The most probable distance observed by DEER is rationalized based on the rotamers observed in the crystal structure. Additionally, the appropriateness of common molecular modeling methods that account for R1 conformational preferences are assessed for the β‐sheet environment. These results show that interpretation of R1 behavior in β‐sheets is difficult and indicate further development is needed for these computational methods to correctly relate DEER distances to protein structure at edge β‐strand sites.  相似文献   

9.
Natural killer T (NKT) cells express a semi-invariant Vα14 T cell receptor (TCR) and recognize structurally diverse antigens presented by the antigen-presenting molecule CD1d that range from phosphoglycerolipids to α- and β-anomeric glycosphingolipids, as well as microbial α-glycosyl diacylglycerolipids. Recently developed antibodies that are specific for the complex of the prototypical invariant NKT (iNKT) cell antigen αGalCer (KRN7000) bound to mouse CD1d have become valuable tools in elucidating the mechanism of antigen loading and presentation. Here, we report the 3.1 Å resolution crystal structure of the Fab of one of these antibodies, L363, bound to mCD1d complexed with the αGalCer analog C20:2, revealing that L363 is an iNKT TCR-like antibody that binds CD1d-presented αGalCer in a manner similar to the TCR. The structure reveals that L363 depends on both the L and H chains for binding to the glycolipid-mCD1d complex, although only the L chain is involved in contacts with the glycolipid antigen. The H chain of L363 features residue Trp-104, which mimics the TCR CDR3α residue Leu-99, which is crucial for CD1d binding. We characterized the antigen-specificity of L363 toward several different glycolipids, demonstrating that whereas the TCR can induce structural changes in both antigen and CD1d to recognize disparate lipid antigens, the antibody L363 can only induce the F′ roof formation in CD1d but fails to reorient the glycolipid headgroup necessary for binding. In summary, L363 is a powerful tool to study mechanism of iNKT cell activation for structural analogs of KRN7000, and our study can aid in the design of antibodies with altered antigen specificity.  相似文献   

10.
Hepatocyte growth factor (HGF) is essential for embryogenesis, tissue regeneration and tumour malignancy through the activation of its receptor, c‐Met. We previously demonstrated that HGF α‐chain hairpin–loop, K1 domain and β‐chain are required for c‐Met signalling. The sequential phosphorylation of tyrosine residues, from c‐Met kinase domain to multidocking regions, is required for HGF‐signalling transduction. Herein, we provide evidence that the disconcerted activation of c‐Met tyrosine regions fails to induce biological functions. When human cells were incubated with ‘mouse HGF’, kinase domain activation (i.e. phospho‐Tyr‐1230/34/35) became evident, but the multidocking site (i.e. Tyr‐1349) was not phosphorylated, resulting in unsuccessful induction of migration and mitogenesis. The binding ability of mouse HGF α‐chain, or of β‐chain, to human c‐Met was lower than that of human HGF, as evidenced by HGF–chimera assay. Notably, only four amino acid positions in HGF α‐chain hairpin–loop and K1 domain and six positions in β‐chain differed between human HGF and mouse HGF. The human‐specific amino acids (such as Gln‐95 in hairpin–loop, Arg‐134 in K1 domain and Cys‐561 in β‐chain) may be important for accurate c‐Met assembly and signalling transduction. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
The small heat shock protein (sHSP) from Methanococcus jannaschii (Mj Hsp16.5) forms a monodisperse 24mer and each of its monomer contains two flexible N‐ and C‐terminals and a rigid α‐crystallin domain with an extruding β‐strand exchange loop. The minimal α‐crystallin domain with a β‐sandwich fold is conserved in sHSP family, while the presence of the β‐strand exchange loop is divergent. The function of the β‐strand exchange loop and the minimal α‐crystallin domain of Mj Hsp16.5 need further study. In the present study, we constructed two fragment‐deletion mutants of Mj Hsp16.5, one with both the N‐ and C‐terminals deleted (ΔNΔC) and the other with a further deletion of the β‐strand exchange loop (ΔNΔLΔC). ΔNΔC existed as a dimer in solution. In contrast, the minimal α‐crystallin domain ΔNΔLΔC became polydisperse in solution and exhibited more efficient chaperone‐like activities to prevent amorphous aggregation of insulin B chain and fibril formation of the amyloidogenic peptide dansyl‐SSTSAA‐W than the mutant ΔNΔC and the wild type did. The hydrophobic probe binding experiments indicated that ΔNΔLΔC exposed much more hydrophobic surface than ΔNΔC. Our study also demonstrated that Mj Hsp16.5 used different mechanisms for protecting different substrates. Though Mj Hsp16.5 formed stable complexes with substrates when preventing thermal aggregation, no complexes were detected when preventing aggregation under non‐heat‐shock conditions. Proteins 2014; 82:1156–1167. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
While the cis‐acyltransferase modular polyketide synthase assembly lines have largely been structurally dissected, enzymes from within the recently discovered trans‐acyltransferase polyketide synthase assembly lines are just starting to be observed crystallographically. Here we examine the ketoreductase (KR) from the first polyketide synthase module of the bacillaene nonribosomal peptide synthetase/polyketide synthase at 2.35‐Å resolution. This KR naturally reduces both α‐ and β‐keto groups and is the only KR known to do so during the biosynthesis of a polyketide. The isolated KR not only reduced an N‐acetylcysteamine‐bound β‐keto substrate to a D ‐β‐hydroxy product, but also an N‐acetylcysteamine‐bound α‐keto substrate to an L ‐α‐hydroxy product. That the substrates must enter the active site from opposite directions to generate these stereochemistries suggests that the acyl‐phosphopantetheine moiety is capable of accessing very different conformations despite being anchored to a serine residue of a docked acyl carrier protein. The features enabling stereocontrolled α‐ketoreduction may not be extensive since a KR that naturally reduces a β‐keto group within a cis‐acyltransferase polyketide synthase was identified that performs a completely stereoselective reduction of the same α‐keto substrate to generate the D ‐α‐hydroxy product. A sequence analysis of trans‐acyltransferase KRs reveals that a single residue, rather than a three‐residue motif found in cis‐acyltransferase KRs, is predictive of the orientation of the resulting β‐hydroxyl group. Proteins 2014; 82:2067–2077. © 2014 Wiley Periodicals, Inc.  相似文献   

13.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

14.
15.
β‐d ‐glucans from mushroom strains play a major role as biological response modifiers in several clinical disorders. Therefore, a specific assay method is of critical importance to find useful and novel sources of β‐d ‐glucans with anti‐tumor activity. Hybridoma technology was used to raise monoclonal antibodies (Mabs) against extracellular β‐d ‐glucans (EBG) from Pleurotus ostreatus. Two of these hybridoma clones (3F8_3H7 and 1E6_1E8_B3) secreting Mabs against EBG from P. ostreatus were selected and 3F8_3H7 was used to investigate if they are polyol‐responsive Mabs (PR‐Mabs) by using ELlSA‐elution assay. This hybridoma cell line secreted Mab of IgM class, which was purified in a single step by gel filtration chromatography on Sephacryl S‐300HR, which revealed a protein band on native PAGE with Mr of 917 kDa. Specificity studies of Mab 3F8_3H7 revealed that it recognized a common epitope on several β‐d ‐glucans from different basidiomycete strains as determined by indirect ELlSA and Western blotting under native conditions. This Mab exhibited high apparent affinity constant (KApp) for β‐d ‐glucans from several mushroom strains. However, it revealed differential reactivity to some heat‐treated β‐d ‐glucans compared with the native forms suggesting that it binds to a conformation‐sensitive epitope on β‐d ‐glucan molecule. Epitope analysis of Mab 3F8_3H7 and 1E6_1E8_B3 was investigated by additivity index parameter, which revealed that they bound to the same epitope on some β‐d ‐glucans and to different epitopes in other antigens. Therefore, these Mab can be used to assay for β‐d ‐glucans as well as to act as powerful probes to detect conformational changes in these biopolymers. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 32:116–125, 2016  相似文献   

16.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
Our goal was to gain a better understanding of how protein stability can be increased by improving β‐turns. We studied 22 β‐turns in nine proteins with 66–370 residues by replacing other residues with proline and glycine and measuring the stability. These two residues are statistically preferred in some β‐turn positions. We studied: Cold shock protein B (CspB), Histidine‐containing phosphocarrier protein, Ubiquitin, Ribonucleases Sa2, Sa3, T1, and HI, Tryptophan synthetase α‐subunit, and Maltose binding protein. Of the 15 single proline mutations, 11 increased stability (Average = 0.8 ± 0.3; Range = 0.3–1.5 kcal/mol), and the stabilizing effect of double proline mutants was additive. On the basis of this and our previous work, we conclude that proteins can generally be stabilized by replacing nonproline residues with proline residues at the i + 1 position of Type I and II β‐turns and at the i position in Type II β‐turns. Other turn positions can sometimes be used if the φ angle is near ?60° for the residue replaced. It is important that the side chain of the residue replaced is less than 50% buried. Identical substitutions in β‐turns in related proteins give similar results. Proline substitutions increase stability mainly by decreasing the entropy of the denatured state. In contrast, the large, diverse group of proteins considered here had almost no residues in β‐turns that could be replaced by Gly to increase protein stability. Improving β‐turns by substituting Pro residues is a generally useful way of increasing protein stability. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

18.
The β‐N‐acetylhexosaminidase FDL specifically removes the β‐1,2‐GlcNAc residue conjugated to the α‐1,3‐mannose residue of the core structure of insect N‐glycans, playing significant physiological roles in post‐translational modification in the Golgi apparatus. Little is known about its enzymatic properties. We obtained the OfFDL gene from the insect Ostrinia furnacalis by RT‐PCR. The full length cDNA of FDL is 2241 bp carrying an opening reading frame of 1923 bp encoding 640 amino acids. The recombinant protein OfFDL in a soluble and active form was obtained with high purity through a two‐step purification strategy. The recombinant OfFDL exclusively hydrolyzes the terminal β‐1,2‐GlcNAc residue from the α‐1,3 branch instead of the α‐1,6 branch of the substrate GnGn‐PA. Several kinetic parameters including kcat/Km values toward four artificial substrates and Ki values of three representative hexosaminidase inhibitors were obtained.  相似文献   

19.
Our previous study showed that intraperitoneal injection of α‐galactosylceramide (α‐GalCer) has the ability to activate lung iNKT cells, but α‐GalCer‐activated iNKT cells do not result in airway inflammation in wild‐type (WT) mice. Many studies showed that iNKT cells had the capacity to induce Treg cells, which gave rise to peripheral tolerance. Therefore, we examined the influence of intraperitoneal administration of α‐GalCer on the expansion and suppressive activity of lung Treg cells using iNKT cell‐knockout mice and co‐culture experiments in vitro. We also compared airway inflammation and airway hyperresponsiveness (AHR) after α‐GalCer administration in specific anti‐CD25 mAb‐treated mice. Our data showed that intraperitoneal injection of α‐GalCer could promote the expansion of lung Treg cells in WT mice, but not in iNKT cell‐knockout mice. However, α‐GalCer administration could not boost suppressive activity of Treg cells in WT mice and iNKT cell‐knockout mice. Interestingly, functional inactivation of Treg cells could induce airway inflammation and AHR in WT mice treated with α‐GalCer. Furthermore, α‐GalCer administration could enhance iNKT cells to secrete IL‐2, and neutralization of IL‐2 reduced the expansion of Treg cells in vivo and in vitro. Thus, intraperitoneal administration of α‐GalCer can induce the generation of lung Treg cells in mice through the release of IL‐2 by the activated iNKT cells.  相似文献   

20.
Secondary structural transitions from α‐helix to β‐sheet conformations are observed in several misfolding diseases including Alzheimer's and Parkinson's. Determining factors contributing favorably to the formation of each of these secondary structures is therefore essential to better understand these disease states. β‐hairpin peptides form basic components of anti‐parallel β‐sheets and are suitable model systems for characterizing the fundamental forces stabilizing β‐sheets in fibrillar structures. In this study, we explore the free energy landscape of the model β‐hairpin peptide GB1 and its E2 isoform that preferentially adopts α‐helical conformations at ambient conditions. Umbrella sampling simulations using all‐atom models and explicit solvent are performed over a large range of end‐to‐end distances. Our results show the strong preference of GB1 and the E2 isoform for β‐hairpin and α‐helical conformations, respectively, consistent with previous studies. We show that the unfolded states of GB1 are largely populated by misfolded β‐hairpin structures which differ from each other in the position of the β‐turn. We discuss the energetic factors contributing favorably to the formation of α‐helix and β‐hairpin conformations in these peptides and highlight the energetic role of hydrogen bonds and non‐bonded interactions. Proteins 2014; 82:2394–2402. © 2014 Wiley Periodicals, Inc.  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