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.
The two molecules of the asymmetric unit of the pH 3.5 conformer of alpha-chymotrypsin have been refined at 1.67-A resolution using restrained least squares methods with Hendrickson's program (PROLSQ). The final R factor is 0.179 (including 247 water molecules). The folding of the main chain of the independent molecules is the same within experimental error but the same does not generally apply to the side chain stereochemistry. From this we conclude that the folding of a protein structure is basically independent of most of the detailed stereochemistry of its side chains. The side chains of the interface region between the independent molecules display pronounced asymmetry. This asymmetry suggests that dynamic and asymmetrical structural changes take place at the time of oligomerization leading to more energetically favorable interactions for the dimer. Comparison of the structures of the independent molecules of alpha-chymotrypsin with the structure of monomeric gamma-chymotrypsin revealed that although the folding of the three molecules is essentially the same, numerous and significant differences pervade the side chain stereochemistry attributable to general flexibility. The specificity site of alpha-chymotrypsin is occupied by ordered water molecules in a similar way to gamma-chymotrypsin and other proteins. Some of these water molecules are displaced when substrate binds to the enzyme, while the others appear to help identify and position the aromatic side chain in catalysis.  相似文献   

2.
A method of determining the water structure in protein crystals is described using neutron solvent difference maps. These maps are obtained by comparing the changes in diffracted intensities between two data sets, one in which H2O is the major solvent constituent, and a second in which D2O is the solvent medium. To a good first approximation, the protein atom contributions to the scattering intensities in both data sets are equal and cancel, but since H2O and D2O have very different neutron-scattering properties, their differences are accentuated to reveal an accurate representation of the solvent structure. The method also employs a series of density modification steps that impose known physical constraints on the density distribution function in the unit cell by making real space modifications directly to the density maps. Important attributes of the method are that (1) it is less subjective in the assignment of water positions than X-ray analysis; (2) there is threefold improvement in the signal-to-noise ratio for the solvent density; and (3) the iterative density modification produces a low-biased representation of the solvent density. Tests showed that water molecules with as low as 10% occupancy could be confidently assigned. About 300 water sites were assigned for trypsin from the refined solvent density; 140 of these sites were defined in the maps as discrete peaks, while the remaining were found within less-ordered channels of density. There is a very good correspondence between the sites in the primary hydration layer and waters found in the X-ray structure. Most water sites are clustered into H-bonding networks, many of which are found along intermolecular contact zones. The bound water is equally distributed between contacting apolar and polar atoms at the protein interface. A common occurrence at hydrophobic surfaces is that apolar atoms are circumvented by one or more waters that are part of a larger water network. When the effects on surface accessibility by neighboring molecules in the crystal lattice are taken into consideration, only about 29% of the surface does not interface ordered water. About 25% of the ordered water is found in the second hydration sphere. In many instances these waters bridge larger clusters of primary layer waters. It is apparent that, in certain regions of the crystal, the organization of ordered water reflects the characteristics of the crystal environment more than those of trypsin's surface alone.  相似文献   

3.
Refined structure of glutathione reductase at 1.54 A resolution   总被引:28,自引:0,他引:28  
The crystal structure of human glutathione reductase has been established at 1.54 A resolution using a restrained least-squares refinement method. Based on 77,690 independent reflections of better than 10 A resolution, a final R-factor of 18.6% was obtained with a model obeying standard geometry within 0.025 A in bond lengths and 2.4 degrees in bond angles. The final 2Fo-Fc electron density map allows for the distinction of carbon, nitrogen and oxygen atoms with temperature factors below about 25 A2. Apart from 461 amino acid residues and the prosthetic group FAD, the model contains 524 solvent molecules, about 118 of which can be considered an integral part of the enzyme. The largest solvent cluster is at the dimer interface and contains 104 interconnected solvent molecules, part of which are organized in a warped sheet-like structure. The main-chain dihedral angles are well-concentrated in the allowed regions of the Ramachandran plot. The spread of dihedral angles in beta-pleated sheets is much larger than in alpha-helices and especially in alpha-helix cores, indicating the higher plasticity of beta-structures. The analysis revealed a large amount of 3(10)-helix. The side-chain conformations cluster at the staggered positions, and show well-defined preferences. Also, a mobility gradient is observed for side-chains. Non-polar and polar side-chains show average temperature factor increases per bond of 10% and 25%, respectively. A number of alternative conformations of internal side-chains, in particular serines and methionines, have been detected. The extended FAD molecule also shows a mobility gradient between the very rigid flavin (mean value of B) = 8.7 A2) and the more mobile adenine (mean value of B = 16.2 A2). The entire active center is particularly well ordered, with temperature factors around 10 A2. The dimer interface consists of a rigid contact area, which is well conserved in the Escherichia coli enzyme, and a flexible area that is not. Altogether, the buried surfaces at the crystal contacts are half as large as at the dimer interface, but less specific. The refined structure shows clearly that there are no buried cations compensating the charge of the pyrophosphate moiety of FAD. The flavin deviates slightly from standard geometry, which is possibly caused by the polypeptide environment. In contrast to an earlier interpretation, atom N5 of the flavin can accommodate a proton, and it is conceivable that this proton proceeds to the redox-active disulfide.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
The solvation of the antibody–antigen Fv D1.3–lysozyme complex is investigated through a study of the conservation of water molecules in crystal structures of the wild-type Fv fragment of antibody D1.3, 5 free lysozyme, the wild-type Fv D1.3–lysozyme complex, 5 Fv D1.3 mutants complexed with lysozyme and the crystal structure of an idiotope (Fv D1.3)-abti-idiotope (Fv E5.2) complex. In all, there are 99 water molecules common to the wild-type and mutant antibody–lysozyme complexes. The antibody–lysozyme interface includes 25 well-ordered solvent molecules, conserved among the wild-type and mutant Fv D1.3–lysozyme complexes, which are bound directly or through other water molecules to both antibody and antigen. In addition to contributing hydrogen bonds to the antibody–antigen interaction the solvent molecules fill many interface cavities. Comparison with x-ray crystal structures of free Fv D1.3 and free lysozyme shows that 20 of these conserved interface waters in the complex were bound to one of the free proteins. Uo to 23 additional water molecules are also found in the antibody–antigen interface, however these waters do no bridge antibody and antigen and their temperature factors are much higher than those of the 25 well-ordered waters. Fifteen water molecules are displaced to form the complex, some of which are substituted by hydrophilic protein atoms, and 5 water molecules are added at the antibody–antigen interface with the formation of the complex. While the current crystal models of the D1.3–lysozyme complex do not demonstrate the increase in bound waters found in a physico-chemical study of the interaction at decreased water activities, the 25 well-ordered interface water contribute a net gain of 10 hydrogen bonds to complex stability.  相似文献   

5.
The molecular structure of the complex between bovine pancreatic alpha-chymotrypsin (EC 3.4.4.5) and the third domain of the Kazal-type ovomucoid from Turkey (OMTKY3) has been determined crystallographically by the molecular replacement method. Restrained-parameter least-squares refinement of the molecular model of the complex has led to a conventional agreement factor R of 0.168 for the 19,466 reflections in the 1.8 A (1 A = 0.1 nm) resolution shell [I greater than or equal to sigma (I)]. The reactive site loop of OMTKY3, from Lys13I to Arg21I (I indicates inhibitor), is highly complementary to the surface of alpha-chymotrypsin in the complex. A total of 13 residues on the inhibitor make 113 contacts of less than 4.0 A with 21 residues of the enzyme. A short contact (2.95 A) from O gamma of Ser195 to the carbonyl-carbon atom of the scissile bond between Leu18I and Glu19I is present; in spite of it, this peptide remains planar and undistorted. Analysis of the interactions of the inhibitor with chymotrypsin explains the enhanced specificity that chymotrypsin has for P'3 arginine residues. There is a water-mediated ion pair between the guanidinium group on this residue and the carboxylate of Asp64. Comparison of the structure of the alpha-chymotrypsin portion of this complex with the several structures of alpha and gamma-chymotrypsin in the uncomplexed form shows a high degree of structural equivalence (root-mean-square deviation of the 234 common alpha-carbon atoms averages 0.38 A). Significant differences occur mainly in two regions Lys36 to Phe39 and Ser75 to Lys79. Among the 21 residues that are in contact with the ovomucoid domain, only Phe39 and Tyr146 change their conformations significantly as a result of forming the complex. Comparison of the structure of the OMTKY3 domain in this complex to that of the same inhibitor bound to a serine proteinase from Streptomyces griseus (SGPB) shows a central core of 44 amino acids (the central alpha-helix and flanking small 3-stranded beta-sheet) that have alpha-carbon atoms fitting to within 1.0 A (root-mean-square deviation of 0.45 A) whereas the residues of the reactive-site loop differ in position by up to 1.9 A (C alpha of Leu18I). The ovomucoid domain has a built-in conformational flexibility that allows it to adapt to the active sites of different enzymes. A comparison of the SGPB and alpha-chymotrypsin molecules is made and the water molecules bound at the inhibitor-enzyme interface in both complexes are analysed for similarities and differences.  相似文献   

6.
7.
Secondary structure of alpha-chymotrypsin in water/ethanol was investigated by circular dichroic (CD) spectroscopy. The changes in catalytic activity were discussed in terms of structural changes of the enzyme. Alpha-chymotrypsin formed beta-sheet structure in water/ethanol (50/50 by volume), but it was substantially less active as compared to that in water. At water/ethanol 10/90, alpha-chymotrypsin took on a native-like structure, which gradually changed to beta conformation with concomitant loss of activity. Change of solvent composition from water/ethanol 50/50 to 90/10 or 10/90 by dilution with water or ethanol, respectively, led to partial recovery of native or native-like structure and activity. In water/methanol, alpha-chymotrypsin tended to form stable beta-sheet structure at water/methanol ratios lower than 50/50, but the catalytic activity decreased with time. Change to alpha-helix structure with substantial loss in catalytic activity was observed when alpha-chymotrypsin was dissolved in water/2,2,2-trifluoroethanol with water contents lower than 50%. In water/2,2,2-trifluoroethanol 90/10, alpha-chymotrypsin initially had the CD spectrum of native structure, but it changed with time to that characteristic of beta-sheet structure.  相似文献   

8.
The dynamics and structure of Serratia marcescens endonuclease and its neighboring solvent are investigated by molecular dynamics (MD). Comparisons are made with structural and biochemical experiments. The dimer form is physiologic and functions more processively than the monomer. We previously found a channel formed by connected clusters of waters from the active site to the dimer interface. Here, we show that dimerization clearly changes correlations in the water structure and dynamics in the active site not seen in the monomer. Our results indicate that water at the active sites of the dimer is less affected compared with bulk solvent than in the monomer where it has much slower characteristic relaxation times. Given that water is a required participant in the reaction, this gives a clear advantage to dimerization in the absence of an apparent ability to use both active sites simultaneously.  相似文献   

9.
Chymotrypsin preparations are contaminated by autolysis products and other post-translational products derived from the zymogen. Some prevalent contaminants are difficult to detect with activity assays and molecular weight separation. However, our differential scanning calorimetry (DSC) studies of alpha-chymotrypsin have revealed that gamma-chymotrypsin is a common contaminant in commercial preparations, and the alpha- and gamma-species differ significantly in thermal stability. Thus, DSC analysis provides a sensitive and rapid means to assess the homogeneity of preparations. Moreover, free metal ion binding studies conducted indicate that the alpha- and gamma-species differ substantially in the number of metal binding sites and metal affinity. Therefore, to attempt to repurify a commercial preparation of alpha-chymotrypsin, a resolubilized sample of alpha-chymotrypsin was subjected to immobilized metal (Cu+2) affinity chromatography with pH elution and the fractions were subjected to DSC analysis. The process successfully removed the majority of the contaminating gamma-chymotrypsin.  相似文献   

10.
In order to address the recognition mechanism of the fragments of antibody variable regions, termed Fv, toward their target antigen, an x-ray crystal structure of an anti-hen egg white lysozyme antibody (HyHEL-10) Fv fragment complexed with its cognate antigen, hen egg white lysozyme (HEL), was solved at 2.3 A. The overall structure of the complex is similar to that reported in a previous article dealing with the Fab fragment-HEL complex (PDB ID code,). However, the areas of Fv covered by HEL upon complex formation increased by about 100 A(2) in comparison with the Fab-HEL complex, and two local structural differences were observed in the heavy chain of the variable region (VH). In addition, small but significant local structural changes were observed in the antigen, HEL. The x-ray data permitted the identification of two water molecules between the VH and HEL and six water molecules retained in the interface between the antigen and the light chain complementarity determining regions (CDRs) 2 and 3 (CDR-L2 and CDR-L3). These water molecules bridge the antigen-antibody interface through hydrogen bond formation in the VL-HEL interface. Eleven water molecules were found to complete the imperfect VH-VL interface, suggesting that solvent molecules mediate the stabilization of interaction between variable regions. These results suggest that the unfavorable effect of deletion of constant regions on the antigen-antibody interaction is compensated by an increase in favorable interactions, including structural changes in the antigen-antibody interface and solvent-mediated hydrogen bond formation upon complex formation, which may lead to a minimum decreased affinity of the antibody Fv fragment toward its antigen.  相似文献   

11.
Bueno M  Camacho CJ 《Proteins》2007,69(4):786-792
Some challenging targets in CAPRI (T24/25 and T26) involve binding solvent accessible acidic residues at the core of the binding interface, where they are always found immersed in crystal waters. In fact, Asp and Glu residues are more likely to form part of the hydrogen bond network of their surrounding crystal water molecules than to form a buried salt bridge. Interestingly, many of the crystal waters mediating the intermolecular interactions of the acidic groups are already present in the unbound structure, reinforcing the notion that some water molecules behave as an extension of the protein structure. This is in contrast to acidic groups found in the periphery of the binding interface that form ubiquitous salt bridges that cement the high affinity complex, while at the same time they are exposed to rapidly exchanging water molecules. Because of this, dichotomy implicit solvent scoring functions fail to properly rank these complexes by prioritizing salt bridges rather than water mediated contacts. A detailed analysis of Target 24, for which our group predicted two out of the four successful homology model complex structures, and Target 26 reveal how crystal waters shape the binding cavities of acidic groups prior to binding, in agreement with the theory of anchor residues as mediators of protein recognition.  相似文献   

12.
Characterization of the early stages of peptide aggregation is of fundamental importance in elucidating the mechanism of the formation of deposits associated with amyloid disease. The initial step in the pathway of aggregation of the Abeta-protein, whose monomeric NMR structure is known, was studied through the simulation of the structure and stability of the peptide dimer in aqueous solution. A protocol based on shape complementarity was used to generate an assortment of possible dimer structures. The structures generated based on shape complementarity were evaluated using rapidly computed estimates of the desolvation and electrostatic interaction energies to identify a putative stable dimer structure. The potential of mean force associated with the dimerization of the peptides in aqueous solution was computed for both the hydrophobic and the electrostatic driven forces using umbrella sampling and classical molecular dynamics simulation at constant temperature and pressure with explicit solvent and periodic boundary conditions. The comparison of the two free energy profiles suggests that the structure of the peptide dimer is determined by the favorable desolvation of the hydrophobic residues at the interface. Molecular dynamics trajectories originating from two putative dimer structures indicate that the peptide dimer is stabilized primarily through hydrophobic interactions, while the conformations of the peptide monomers undergo substantial structural reorganization in the dimerization process. The finding that the phi-dimer may constitute the ensemble of stable Abeta(10-35) dimer has important implications for fibril formation. In particular, the expulsion of water molecules at the interface might be a key event, just as in the oligomerization of Abeta(16-22) fragments. We conjecture that events prior to the nucleation process themselves might involve crossing free energy barriers which depend on the peptide-peptide and peptide-water interactions. Consistent with existing experimental studies, the peptides within the ensemble of aggregated states show no signs of formation of secondary structure.  相似文献   

13.
To investigate the stability and functional role of long-residence water molecules in the Q61H variant of the signaling protein K-ras, we analyzed all available Ras crystal structures and conformers derived from a series of independent explicit solvent molecular dynamics (MD) simulations totaling 1.76 μs. We show that the protein samples a different region of phase space in the presence and absence of several crystallographically conserved and buried water molecules. The dynamics of these waters is coupled with the local as well as the global motions of the protein, in contrast to less buried waters whose exchange with bulk is only loosely coupled with the motion of loops in their vicinity. Aided by two novel reaction coordinates involving the distance (d) between the C(α) atoms of G60 at switch 2 and G10 at the P-loop and the N-C(α)-C-O dihedral (ξ) of G60, we further show that three water molecules located in lobe1, at the interface between the lobes and at lobe2, are involved in the relative motion of residues at the two lobes of Q61H K-ras. Moreover, a d/ξ plot classifies the available Ras x-ray structures and MD-derived K-ras conformers into active GTP-, intermediate GTP-, inactive GDP-bound, and nucleotide-free conformational states. The population of these states and the transition between them is modulated by water-mediated correlated motions involving the functionally critical switch 2, P-loop and helix 3. These results suggest that water molecules act as allosteric ligands to induce a population shift among distinct switch 2 conformations that differ in effector recognition.  相似文献   

14.
M Harel  C T Su  F Frolow  I Silman  J L Sussman 《Biochemistry》1991,30(21):5217-5225
The determination of three separate gamma-chymotrypsin structures at different temperatures and resolutions confirmed the presence of electron density in the active site, which could be interpreted as an oligopeptide as had previously been suggested by Dixon and Matthews [(1989) Biochemistry 28, 7033-7038]. HPLC analyses of the enzyme before and after crystallization demonstrated the presence of a wide variety of oligopeptides in the redissolved crystal, most with COOH-terminal aromatic residues, as expected of the products of chymotrypsin cleavage, which appeared to arise from extensive autolysis of the enzyme under the crystallization conditions. The refined structures agree well with the conformation of both gamma-chymotrypsin and alpha-chymotrypsin. The electron density in the active site is thus interpreted as arising from a repertoire of autolysed oligopeptides produced concomitantly with crystallization. The COOH-terminal carbons of the polypeptide(s) display short contact distances (1.97, 2.47, and 2.13 A, respectively) to Ser195 O gamma in all three refined structures, but the electron density is not continuous between these two atoms in any of them. This suggests that some sequences are covalently bound as enzyme intermediates while others are noncovalently bound as enzyme-product complexes.  相似文献   

15.
A new method to analyze the distribution of water molecules around the bases in DNA is presented. This method relies on the notion of a "hydrated building block," which represents the joint observed hydration around all bases of a particular type, in structures of a particular conformation type. The hydrated building blocks were constructed using atomic coordinates from 40 structures contained in the Nucleic Acid Database. Pseudoelectron densities were calculated for water molecules in each hydrated building block using standard crystallographic procedures. The electron densities were fitted to obtain "average building blocks," which represent bases with waters only at average or probable positions. Both types of building blocks were used to construct models of hydrated DNA oligomers. The essential features of the solvent structure around d(CGCGAATTCGCG)2 in the B form and d(CGCGCG)2 in the Z form were reproduced.  相似文献   

16.
Reiling KK  Pray TR  Craik CS  Stroud RM 《Biochemistry》2000,39(42):12796-12803
The structure of Kaposi's sarcoma-associated herpesvirus protease (KSHV Pr), at 2.2 A resolution, reveals the active-site geometry and defines multiple possible target sites for drug design against a human cancer-producing virus. The catalytic triad of KSHV Pr, (Ser114, His46, and His157) and transition-state stabilization site are arranged as in other structurally characterized herpesviral proteases. The distal histidine-histidine hydrogen bond is solvent accessible, unlike the situation in other classes of serine proteases. As in all herpesviral proteases, the enzyme is active only as a weakly associated dimer (K(d) approximately 2 microM), and inactive as a monomer. Therefore, both the active site and dimer interface are potential targets for antiviral drug design. The dimer interface in KSHV Pr is compared with the interface of other herpesviral proteases. Two conserved arginines (Arg209), one from each monomer, are buried within the same region of the dimer interface. We propose that this conserved arginine may provide a destabilizing element contributing to the tuned micromolar dissociation of herpesviral protease dimers.  相似文献   

17.
Binding of the estrogen receptor to DNA. The role of waters.   总被引:2,自引:0,他引:2  
Molecular dynamics simulations are carried out to investigate the binding of the estrogen receptor, a member of the nuclear hormone receptor family, to specific and non-specific DNA. Two systems have been simulated, each based on the crystallographic structure of a complex of a dimer of the estrogen receptor DNA binding domain with DNA. One structure includes the dimer and a consensus segment of DNA, ds(CCAGGTCACAGTGACCTGG); the other structure includes the dimer and a nonconsensus segment of DNA, ds(CCAGAACACAGTGACCTGG). The simulations involve an atomic model of the protein-DNA complex, counterions, and a sphere of explicit water with a radius of 45 A. The molecular dynamics package NAMD was used to obtain 100 ps of dynamics for each system with complete long-range electrostatic interactions. Analysis of the simulations revealed differences in the protein-DNA interactions for consensus and nonconsensus sequences, a bending and unwinding of the DNA, a slight rearrangement of several amino acid side chains, and inclusion of water molecules at the protein-DNA interface region. Our results indicate that binding specificity and stability is conferred by a network of direct and water mediated protein-DNA hydrogen bonds. For the consensus sequence, the network involves three water molecules, residues Glu-25, Lys-28, Lys-32, Arg-33, and bases of the DNA. The binding differs for the nonconsensus DNA sequence in which case the fluctuating network of hydrogen bonds allows water molecules to enter the protein-DNA interface. We conclude that water plays a role in furnishing DNA binding specificity to nuclear hormone receptors.  相似文献   

18.
The three-dimensional crystal structure of the (R207S, R292S) mutant of malate dehydrogenase from Haloarcula marismortui was solved at 1.95A resolution in order to determine the role of salt bridges and solvent ions in halophilic adaptation and quaternary structure stability. The mutations, located at the dimer-dimer interface, disrupt two inter-dimeric salt bridge clusters that are essential for wild-type tetramer stabilisation. Previous experiments in solution, performed on the double mutant, had shown a tetrameric structure in 4M NaCl, which dissociated into active dimers in 2M NaCl. In order to establish if the active dimeric form is a product of the mutation, or if it also exists in the wild-type protein, complementary studies were performed on the wild-type enzyme by analytical centrifugation and small angle neutron scattering experiments. They showed the existence of active dimers in NaF, KF, Na(2)SO(4), even in the absence of NADH, and in the presence of NADH at concentrations of NaCl below 0.3M. The crystal structure shows a tetramer that, in the absence of the salt bridge clusters, appears to be stabilized by a network of ordered water molecules and by Cl(-) binding at the dimer-dimer interface. The double mutant and wild-type dimer folds are essentially identical (the r.m.s. deviation between equivalent C(alpha) positions is 0.39A). Chloride ions are also observed at the monomer-monomer interfaces of the mutant, contributing to the stability of each dimer against low salt dissociation. Our results support the hypothesis that extensive binding of water and salt is an important feature of adaptation to a halophilic environment.  相似文献   

19.
The solvent structure in orthorhombic crystals of bovine trypsin has been independently determined by X-ray diffraction to 1.35 A resolution and by neutron diffraction to 2.1 A resolution. A consensus model of the water molecule positions was obtained using oxygen positions identified in the electron density map determined by X-ray diffraction, which were verified by comparison to D2O-H2O difference neutron scattering density. Six of 184 water molecules in the X-ray structure, all with B-factors greater than 50 A2, were found to be spurious after comparison with neutron results. Roughly two-thirds of the water of hydration expected from thermodynamic data for proteins was localized by neutron diffraction; approximately one-half of the water of hydration was located by X-ray diffraction. Polar regions of the protein are well hydrated, and significant D2O-H2O difference density is seen for a small number of water molecules in a second shell of hydration. Hydrogen bond lengths and angles calculated from unconstrained refinement of water positions are distributed about values typically seen in small molecule structures. Solvent models found in seven other bovine trypsin and trypsinogen and rat trypsin structures determined by X-ray diffraction were compared. Internal water molecules are well conserved in all trypsin structures including anionic rat trypsin, which is 65% homologous to bovine trypsin. Of the 22 conserved waters in trypsin, 19 were also found in trypsinogen, suggesting that they are located in regions of the apoprotein that are structurally conserved in the transition to the mature protein. Seven waters were displaced upon activation of trypsinogen. Water structure at crystal contacts is not generally conserved in different crystal forms. Three groups of integral structural water molecules are highly conserved in all solvent structures, including a spline of water molecules inserted between two beta-strands, which may resemble an intermediate in the formation of beta sheets during the folding of a protein.  相似文献   

20.
Structure of alpha-chymotrypsin refined at 1.68 A resolution   总被引:16,自引:0,他引:16  
  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