首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Human C4b-binding protein (C4BP) functions as a cofactor to factor I in the degradation of C4b and accelerates the decay rate of the C4b2a complex. In this study we describe a monoclonal antibody directed against the alpha-chain of C4BP that inhibits the binding of C4b to C4BP. In order to identify the structural domain of the alpha-chain of C4BP that interacts with C4b, tryptic fragments of C4BP were generated. Amino acid sequence analysis of the fragments revealed that the residues Ser333-Arg356 of the alpha-chain of C4BP contain the epitope of this antibody, and as a consequence, that this part of the alpha-chain of C4BP is likely to be involved in the interaction with C4b.  相似文献   

2.
Two functionally distinct but homologous sites in complement receptor type 1 (CR1) (CD35) were further characterized by homologous substitution mutagenesis of two CR1 derivatives, each containing one site. In both sites, reducing negative and/or increasing positive charge augmented interaction with iC3/C3b and C4b, supporting a role of ionic forces in the binding reaction. In one case, substitution of Asp at the end of complement control protein repeat (CCP) 2 with an Asn transformed the protein, with negligible cofactor activity and iC3 binding, into a mutant with activities similar to native CR1. Consequently, this protein, one-fourth the size of CR1, is a therapeutic candidate for a complement inhibitor. Another important observation is that the residues between two CCPs contribute to activity, probably because they influence positioning of one CCP relative to the next. The initial characterization of the third CCP of an active site led to identification of three peptides necessary for binding. In line with earlier findings for the first two CCPs, interactions with iC3/C3b are similar but not identical to those with C4b, implying overlapping but distinct binding domains. Moreover, changes in cofactor activity usually, but not always, parallel alterations in binding, indicating that these two activities are separable. We also mapped epitopes for a blocking and a function enhancing monoclonal antibody. Their effects can be explained by epitope location. The first antibody binds near functionally important residues. The second may shield inhibitory (negatively charged) residues. These results represent a comprehensive analysis of the active sites of CR1, which is built of modules found in more than 50 mammalian proteins.  相似文献   

3.
There is considerable variation in the TCR repertoire diversity selected by different peptide Ags. Certain responses show limited V region bias with minimal restrictions in the remainder of the sequence while others can be dominated by a single TCR clonotype repeatedly isolated from different individuals. CTL specific for a Kb-restricted determinant from the herpes simplex virus glycoprotein B (gB) preferentially express a dominant TCRBV10 beta-chain subset with extensive conservation located at the V-D junction. However, unlike some biased responses, no single beta-chain V-D-J combination appears to dominate these CTL. Different animals respond with a large array of unique or "private" beta-chain sequences with little J region preference. Here we examine the contribution of the TCR alpha-chain to the gB-specific CTL diversity. The TCR alpha-chains from different TCRBV10-positive gB-specific CTL clones were found to exhibit extensive sequence variation. However, when T cells were forced to use a single alpha-chain in TCR alpha-chain transgenic mice, gB-specific CTL showed limited variation in their beta-chain selection. These T cells retained the TCRBV10 bias but were now dominated by a single beta-chain sequence that could be repeatedly isolated from different transgenic animals. This "public" TCR consisted of the transgenic alpha-chain and a common TCRBV10D2J2S6 beta-chain. These results suggest that preferential use of one TCR subunit can restrict the level of diversity in the other chain due to interchain interactions involving J-derived sequences.  相似文献   

4.
The TCR found on CD4 T cells recognizes peptides bound to self MHC class II molecules as well as non-self MHC class II molecules. We have used the receptor on a cloned T cell line called D10.G4.1 (D10) to perform a structure-function analysis of this interaction. The D10 T cell clone recognizes not only a peptide from conalbumin (CA-wt) bound to syngeneic I-Ak against which it was raised, but also the allogeneic MHC molecules I-A(b,v,p,q,d). In the present study, we show that residue 30 in complementarity-determining region 1 (CDR1) of the TCR alpha-chain interacts with the I-A alpha-chain at hvr2 (residues 52, 53, and 55). We also show that residue 51 in CDR2 of the TCR alpha-chain interacts with the peptide at peptide residue 2. Finally, we show that residue 29 in CDR1 of the TCR beta-chain affects recognition of the glutamic acid at residue 66 in the I-A beta-chain. These data suggest an orientation of TCR relative to its peptide:MHC class II ligands. We argue that this orientation will be shared by all CD4 TCRs, and that it is only subtly different from the common orientation proposed for receptors binding to MHC class I.  相似文献   

5.
C4BP (C4b-binding protein) is a high-molecular-weight plasma protein that inhibits the classical pathway of complement activation. Recent experiments have demonstrated that C4BP binds to many strains of the gram-positive bacterium Streptococcus pyogenes, a major respiratory tract pathogen. Binding to S. pyogenes was shown to be due to members of the M protein family, a group of surface proteins important for virulence. Here we report that human C4BP also binds to all clinical isolates of the gram-negative bacterium Bordetella pertussis, the etiologic agent of whooping cough. In addition, binding of C4BP was demonstrated for other Bordetella species that can cause disease in humans. Characterization of different B. pertussis mutants showed that the binding of C4BP is strongly dependent on the expression of the cell surface protein filamentous hemagglutinin, a well-known virulence factor. Inhibition experiments suggested that B. pertussis and S. pyogenes bind to the same region in C4BP. The finding that B. pertussis and S. pyogenes both have the ability to bind human C4BP suggests that these two unrelated respiratory tract pathogens may use a common mechanism during the establishment of an infection.  相似文献   

6.
Hepatocyte growth factor (HGF) is a heterodimeric molecule composed of the alpha-chain containing the N-terminal hairpin domain, four kringle domains, and the serine protease-like beta-chain. We prepared HGF/NK4 and HGF/beta from the entire HGF after single-cut digestion with elastase. HGF/NK4 contains the N-terminal hairpin and four kringle domains, while HGF/beta is composed of the C-terminal 16 amino acids of the alpha-chain and the entire beta-chain, linked by a disulfide bridge. HGF/NK4 competitively inhibited the binding of 125I-HGF to the receptor, and affinity cross-linking analysis indicated that HGF/NK4 alone can bind to the c-Met receptor. In contrast, HGF/beta alone did not competitively inhibit the binding of 125I-HGF to the receptor and did not bind to the c-Met/HGF receptor. Scatchard analysis and affinity cross-linking experiments indicated that HGF/beta specifically binds to c-Met in the presence of HGF/NK4 but not HGF/NK2. Neither HGF/NK4 nor HGF/beta alone induced mitogenic, motogenic (cell scattering), and morphogenic (induction of branching tubulogenesis) responses; however, HGF/beta did induce these biological responses in the presence of HGF/NK4. Consistent with these results, although neither HGF/NK4 alone nor HGF/beta alone induced tyrosine phosphorylation of the c-Met/HGF receptor, HGF/beta induced tyrosine phosphorylation of the receptor when c-Met/HGF receptor was occupied by HGF/NK4. These results indicate that HGF/beta binds to the c-Met/HGF receptor that is occupied by HGF/NK4 and induces receptor tyrosine phosphorylation and the subsequent biological activities of HGF. We propose that there exists a unique cooperative interaction between alpha- and beta-chains, this interaction leading to beta-chain-dependent receptor tyrosine phosphorylation and subsequent biological responses.  相似文献   

7.
In all other species analyzed to date, the functionally active form of complement component C3 exists as the product of a single gene. We have now identified and characterized three functional C3 proteins (C3-1, C3-3, and C3-4) in trout that are the products of at least two distinct C3 genes. All three proteins are composed of an alpha-and a beta-chain and contain a thioester bond in the alpha-chain. However, they differ in their electrophoretic mobility, glycosylation, reactivity with monospecific C3 antibodies, and relative ability to bind to various surfaces (zymosan, Escherichia coli, erythrocytes). A comparison of the partial amino acid sequences of the three proteins showed that the amino acid sequence identity/similarity of C3-3 to C3-4 is 87/91%, while that of C3-3 and C3-4 to C3-1 is 51.5/65.5% and 60/73% respectively. Thus, trout possess multiple forms of functional C3 that represent the products of several distinct genes and differ in their ability to bind covalently to various complement activators.  相似文献   

8.
The bacterial superantigen staphylococcal enterotoxin A (SEA) binds with high affinity to major histocompatibility complex (MHC) class II molecules and subsequently activates T cells bearing particular T-cell receptor (TCR) Vbeta chains. Structural and mutational studies have defined two distinct MHC class II binding sites located in the N-terminal and C-terminal domains of SEA. The N-terminal F47 amino acid is critically involved in a low-affinity interaction to the MHC class II alpha-chain, while the C-terminal residues H187, H225, and D227 coordinate a Zn2+ ion and bind with moderate affinity to the beta-chain. In order to analyze whether the SEA-MHC class II alpha-chain interaction plays a role in dictating the in vivo repertoire of T-cell subsets, we studied distinct Vbeta populations after stimulation with wild-type SEA [SEA(wt)] and SEA with an F47A mutation [SEA(F47A)]. Injections of SEA(wt) in C57BL/6 mice induced cytokine release in serum, strong cytotoxic T-lymphocyte activity, expansion of T-cell subsets, and modulated expression of the T-cell activation antigens CD25, CD11a, CD44, CD62L, and CD69. SEA-reactive TCR Vbeta3+ and Vbeta11+ T cells were activated, while TCR Vbeta8+ T cells remained unaffected. The SEA(F47A) mutant protein induced a weaker T-cell response and failed to induce substantial interleukin-6 production compared to SEA(wt). Notably, SEA(F47A) failed to activate TCR Vbeta11+ T cells, whereas in vivo expansion and modulation of T-cell activation markers on TCR Vbeta3+ T cells were similar to those for SEA(wt). A similar response to SEA(F47A) was seen among CD4+ and CD8+ T cells. Activation of TCR Vbeta3+ and TCR Vbeta11+ T-cell hybridomas confirmed that SEA(F47A) activates TCR Vbeta3+ but not TCR Vbeta11+ T cells. The data support the view that the SEA-N-terminal MHC class II alpha-chain interaction defines a topology that is required for engagement of certain TCR Vbeta chains in vivo.  相似文献   

9.
Haptoglobin binding to haemoglobin and its isolated alpha- and beta-chains was studied by use of a highly sensitive solid-phase radiometric assay. As expected, adsorbents of haemoglobin bound 125I-labelled haptoglobin more efficiently than did adsorbents of the alpha-chain. However, unexpectedly, adsorbents of the beta-chain were found to be essentially identical with those of the alpha-chain in their ability to bind haptoglobin. These results demonstrate, unequivocally, the ability of beta-chains to bind to haptoglobin, and indicate that this assay is particularly convenient and useful for studying haptoglobin interactions with haemoglobin and its alpha- and beta-chains.  相似文献   

10.
Human protein S binds to C4b-binding protein (C4BP) both in plasma and in a system using purified proteins. Amino acid residues 420-434 of the first disulfide loop of the sex hormone binding globulinlike domain of protein S are involved in the interaction of protein S with C4BP. To define the involvement of specific polar amino acids within residues 420-434, we studied in parallel synthetic protein S peptides and recombinant protein S variants containing the same amino acid replacements, K423E, E424K, Q427E and K429E. Synthetic peptide analogs of peptide PSP-420 (residues 420-434) were assayed for binding C4BP and as inhibitors of complex formation. The PSP-420 peptide and the analogous peptide with the substitution E424K, but not the peptides containing the substitutions K423E and K429E, were able to bind C4BP. Recombinant proteins with mutations of K423E, Q427E and K429E showed reduced affinity for C4BP compared to plasma protein S, recombinant wild type protein S, or E424K-protein S. These results suggest that Lys-423, Gln-427 and Lys-429 of protein S are important for normal binding to C4BP. The anti-protein S monoclonal antibody LJ-56, raised against peptide PSP-420, recognizes only free protein S and inhibits complex formation with C4BP. Antibody LJ-56 recognized the E424K and Q427E peptides but not the K423E or K429E peptides. Similarly, the E424K and Q427E protein S mutants were recognized by LJ-56, whereas the K423E and K429E protein S mutants were not recognized. This suggests that both in the peptide PSP-420 and in protein S, Lys-423 and Lys-429 significantly contribute to binding to antibody LJ-56. These results demonstrate that protein S residues 423, 427 and 429, but not residue 424, are involved in binding to both the antibody LJ-56 and to C4BP. When peptides PSP 420 and SL-6 (residues 447-460) with carboxyterminal amide or carboxylate moieties were compared to their ability to inhibit C4BP-protein S complexation, PSP-420-amide was the most potent. This finding together with the other results described here supports the hypothesis that the residues 420 and 434 in protein S provides a major binding site for C4BP.  相似文献   

11.
Horse alpha-chain inhibits sickle beta-chain-dependent polymerization; however, its inhibitory potential is not as high as that of mouse alpha-chain. Horse alpha-(1-30) and alpha-(31-141) segments make, respectively, minor and major contributions to the inhibitory potential of horse alpha-chain. The sum of the inhibitory potential of the two segments does not account for the inhibitory potential of the full-length horse alpha-chain. Although the polymerization inhibitory potential of horse alpha-chain is lower than mouse alpha-chain, the inhibitory potential of horse alpha-(31-141) is comparable to that of mouse alpha-(31-141). When mouse alpha-(1-30) is stitched to horse alpha-(31-141), the product is a chimeric alpha-chain with an inhibitory potential greater than mouse alpha-chain. In contrast, the stitching of horse alpha-(1-30) with mouse alpha-(31-141) had no additional inhibitory potential. Molecular modeling studies of HbS containing the mouse-horse chimeric alpha-chain indicate altered side-chain interactions at the alpha1beta1 interface when compared with HbS. In addition, the AB/GH corner perturbations facilitate a different stereochemistry for the interaction of the epsilon-amino group of Lys-16(alpha) with the beta-carboxyl group of Asp-116(alpha), resulting in a decrease in the accessibility of the side chain of Lys-16(alpha) to the solvent. Based on molecular modeling, we speculate that these perturbations by themselves, or in synergy with the altered conformational aspects of the alpha1beta1 interactions, represent the molecular basis of the superinhibitory potential of the mouse-horse chimeric alpha-chains.  相似文献   

12.
Various common C4 gene products were isolated from serum by immunoprecipitation. After reduction the C4 alpha-, beta-, and gamma-polypeptide chains were studied by two-dimensional electrophoresis. Isoelectrofocusing was performed in the first dimension and sodium dodecyl sulphate polyacrylamide gradient gel electrophoresis in the second. The charge differences behind the electrophoretic C4 polymorphism were shown to reside in the 95,000-u (atomic mass units) alpha-chain. Charge variation closely mirroring the alpha-chain differences were also found in a 49,000-u fragment of the alpha-chain, most probably C4d. The basic beta-chain could not be studied in detail, but no differences were observed with regard to molecular weight or charge of the gamma-chains of the different C4 gene products.  相似文献   

13.
The glycosylation of murine C4 (Ss protein) synthesized by peritoneal macrophages has been investigated. Both the intracellular precursor, P-C4 (185), and the C4 alpha- and beta-chains that were secreted into medium were found to be glycosylated; however, no carbohydrate units were detected on the gamma-chain. Analyses of the oligosaccharide units showed that P-C4 (185) appears to contain both a "complex" and a "high mannose" carbohydrate group, the alpha-chain a "complex" group and the beta-chain a "high mannose" carbohydrate unit.  相似文献   

14.
To assess the relationship between capillary leakage and inflammatory mediators during sepsis, blood samples were taken on hospital admission, as well as 24 and 72 h later, from 52 children (median age, 3.3 years) with severe meningococcal sepsis, of whom 38 survived and 14 died. Parameters related to cytokines (interleukin 6 [IL-6] IL-8, plasma phospholipase A2, and C-reactive protein [CRP]), to neutrophil degranulation (elastase and lactoferrin), to complement activation (C3a, C3b/c, C4b/c, and C3- and C4-CRP complexes), and to complement regulation (functional and inactivated C1 inhibitor and C4BP) were determined. The degree of capillary leakage was derived from the amount of plasma infused and the severity of disease by assessing the pediatric risk of mortality (PRISM) score. Levels of IL-6, IL-8, C3b/c, C3-CRP complexes, and C4BP on admission, adjusted for the duration of skin lesions, were significantly different in survivors and nonsurvivors (C3b/c levels were on average 2.2 times higher in nonsurvivors, and C3-CRP levels were 1.9 times higher in survivors). Mortality was independently related to the levels of C3b/c and C3-CRP complexes. In agreement with this, levels of complement activation products correlated well with the PRISM score or capillary leakage. Thus, these data show that complement activation in patients with severe meningococcal sepsis is associated with a poor outcome and a more severe disease course. Further studies should reveal whether complement activation may be a target for therapeutical intervention in this disease.  相似文献   

15.
Many diverse effects of heparin on the complement system have been reported. In only a few cases have the sites or the mechanisms of these effects been identified. In order to understand these results we sought to comprehensively analyze which complement proteins interact with heparin and which do not. Purified components of the classical, alternative and terminal pathways of complement were radiolabeled and their affinity for heparin determined. Affinity chromatography of normal human serum on heparin-agarose allowed a complete analysis of complement proteins and confirmed the results obtained with radiolabeled purified components. Of the 22 complement proteins examined, 13 bound heparin (C1q, C2, C4, C4bp, C1INH, B, D, H, P, C6, C8, C9, and vitronectin) while 9 did not bind heparin (C1r, C1s, C3, Factor I, C5, C7, C3b, Ba and Bb). These observations help explain the many effects heparin has on the complement system and they identify the proteins which need to be examined in order to explain these effects.  相似文献   

16.
Vesicular stomatitis virus (VSV) elicits H-2Kb-restricted CTLs specific for the immunodominant VSV octapeptide RGYVYQGL. To study the structural features important for interaction between the TCR beta-chain and the peptide/MHC complex, we immunized TCR alpha-chain transgenic mice with the VSV peptide and raised a panel of anti-VSV CTL clones with identical TCR alpha-chains. Consistent with our previous analysis of uncloned populations of primary CTLs, the anti-VSV CTL clones were all Vbeta13+ and expressed TCR beta-chains with highly homologous complementarity-determining region 3 (CDR3) loops. Although the clones expressed similar TCRs, they differed in their ability to cross-react with VSV peptide variants singly substituted at TCR contact positions 4 and 6. These findings allowed us to identify short stretches of amino acids in the C-terminal region of the CDR3beta loop that, when altered, modify the cross-reaction capability of the TCR to position 4 and position 6 variant peptides. To further probe the structural correlates of biologic cross-reactivity, we used cross-reactive CTL clones and cell lines expressing point mutations in H-2Kb to investigate the effect of single amino acid changes in the peptide on the pattern of recognition of the TCR for the peptide/MHC complex. Single conservative substitutions in the peptide were sufficient to alter the recognition contacts between a cross-reactive TCR and the MHC molecule, supporting the idea that the TCR can make overall structural adjustments in MHC contacts to accommodate single amino acid changes in the peptide.  相似文献   

17.
Tay-Sachs disease is a severe neurodegenerative disorder due to mutations in the HEXA gene coding for the alpha-chain of the alpha-beta heterodimeric lysosomal enzyme beta-hexosaminidase A (HexA). Because no treatment is available for this disease, we have investigated the possibility of enzymatic correction of HexA-deficient cells by HEXA gene transfer. Human HEXA cDNA was subcloned into a retroviral plasmid generating to G.HEXA vector. The best Psi-CRIP producer clone of G.HEXA retroviral particles was isolated, and murine HexA-deficient fibroblasts derived from hexa -/- mice were transduced with the G.HEXA vector. Transduced cells overexpressed the alpha-chain, resulting in the synthesis of interspecific HexA (human alpha-chain/murine beta-chain) and in a total correction of HexA deficiency. The alpha-chain was secreted in the culture medium and taken up by HexA-deficient cells via mannose-6-phosphate receptor binding, allowing for the restoration of intracellular HexA activity in non-transduced cells.  相似文献   

18.
Several extracellular modular proteins, including proteases of the complement and blood coagulation cascades, are shown here to exhibit conserved sequence patterns specific for a particular module-domain association. This was detected by comparative analysis of sequence variability in different multiple sequence alignments, which provides a new tool to investigate the evolution of modular proteins. A first example deals with the proteins featuring a common complement control protein (CCP) module-serine protease (SP) domain pattern at their C-terminal end, defined here as the CCP-SP sub-family. These proteins include the complement proteases C1r, C1s and MASPs, the Limulus clotting factor C, and the proteins of the haptoglobin family. A second example deals with blood coagulation factors VII, IX and X and protein C, all featuring a common epidermal growth factor (EGF)-SP C-terminal assembly. Highly specific motifs are found at the connection between the CCP or EGF module and the activation peptide of the SP domain: [P/A]-x-C-x-[P/A]-[I/V]-C-G-x-[P/S/K] in the case of the CCP-SP proteins, and C-x-[P/S]-x-x-x-[Y/F]-P-C-G in the case of the EGF-SP proteins. Each motif is strictly conserved in the whole sub-family and it is detected in no more than one other known protein sequence. Strikingly, most of the conserved residues specific to each sub-family appear to be clustered at the interface between the SP domain and the CCP or EGF module. We propose that a rigid module-domain interaction occurs in these proteins and has been conserved through evolution. The functional implications of these assemblies, underlined by such evolutionary constraints, are discussed.  相似文献   

19.
Mapping approaches employing blocking antibodies and synthetic peptides have implicated the 727-767 segment at the NH2 terminus of C3b alpha'-chain as contributing to the interactions with factor B, factor H, and CR1. Our previous mutagenesis study on the NH2-terminal acidic cluster of this segment identified residues Glu-736 and Glu-737 as contributing to the binding of C3b to factor B and CR1 but not factor H. We have now extended the charged residue mutagenic scan to cover the remainder of the segment (738-767) and have assessed the ability of the C3b-like C3(H2O) form of the mutant molecules to interact with factor H, CR1, and membrane cofactor protein (MCP) using a cofactor-dependent factor I cleavage assay as a surrogate binding assay. We have found that the negatively charged side chains of Glu-744 and Glu-747 are important for the interaction between C3(H2O) and factor H, a result in general agreement with an earlier synthetic peptide study (Fishelson, Z. (1991) Mol. Immunol. 28, 545-552) which implicated residues within the 744-754 segment in H binding. The interactions of the mutants with soluble CR1 (sCR1) revealed two classes of residues. The first are residues required for sCR1 to be an I cofactor for the first two cleavages of alpha-chain. These are all acidic residues and include the Glu-736/Glu-737 pair, Glu-747, and the Glu-754/Asp-755 pairing. The second class affects only the ability of sCR1 to be a cofactor for the third factor I cleavage and include Glu-744 and the Lys-757/Glu-758 pairing. The dominance of acidic residues in the loss-of-function mutants is striking and suggests that H and CR1 contribute basic residues to the interface. Additionally, although there is partial overlap, the contacts required for CR1 binding appear to extend over a wider portion of the 727-767 segment than is the case for factor H. Finally, none of the mutations had any effect on the interaction between soluble MCP and C3(H2O), indicating that despite its functional homology to H and CR1, MCP differs in its mode of binding to C3b/C3(H2O).  相似文献   

20.
A combination of intrinsic fluorescence and circular dichroic (CD) spectroscopy has been used to characterize the complexes formed between bovine retinal arrestin and heparin or phytic acid, two ligands that are known to mimic the structural changes in arrestin attending receptor binding. No changes in the CD spectra were observed upon ligand binding, nor did the degree of tryptophan fluorescence quenching change significantly in the complexes. These data argue against any large-scale changes in protein secondary or tertiary structure accompanying ligand binding. The change in tyrosine fluorescence intensity was used to determine the dissociation constants for the heparin and phytic acid complexes of arrestin. The only change observed was a saturable diminution of tyrosine fluorescence signal from the protein. For both ligands, the data suggest two distinct binding interactions with the protein--a high--affinity interaction with Kd between 200 and 300 nM, and a lower affinity interaction with Kd between 2 and 8 microM. Study of collisional quenching of tyrosine fluorescence in free arrestin and the ligand-replete complexes indicates that 10 of the 14 tyrosine residues of the protein are solvent-exposed in the free protein; this value drops to between 5 and 6 solvent-exposed residues in the high-affinity complexes of the two ligands. These data suggest that ligand binding leads to direct occlusion of between 4 and 5 tyrosine residues on the solvent-exposed surface of the protein, but not to any large-scale changes in protein structure. The large activation energy previously reported to be associated with arrestin-receptor interactions may therefore reflect localized movements of the N- and C-termini of arrestin, which are proposed to interact in the free protein through electrostatic interactions. Binding of the anionic ligands heparin, phytic acid, or phosphorylated rhodopsin may compete with the C-terminus of arrestin for these electrostatic interactions, thus allowing the C-terminus to swing out of the binding region.  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