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.
Sedimentation equilibrium experiments indicate that neither human chymotrypsin II nor bovine δ-chymotrypsin molecules undergo association in the pH range 3–5 where dimerization occurs with α-chymotrypsin. The weight-average molecular weights of human chymotrypsin II and δ-chymotrypsin in a pH 4.4 0.1 ionic strength buffer are 26,200 and 26,400, respectively, using the measured partial specific volumes of 0.722 and 0.727 ml/g at 25 °C. Number-average molecular weight calculations also support the presence of monomeric species at this pH. In the pH range 6–7.6 where sedimentation velocity studies have shown that δ-chymotrypsin associates at concentrations above 3 mg/ml, no association was observed for either the human chymotrypsin II or bovine δ-chymotrypsin in the sedimentation equilibrium experiments where protein concentrations were below 1.2 mg/ml. These studies provide additional evidence that human chymotrypsin II is similar to bovine δ-chymotrypsin.  相似文献   

2.
The molecular weight of delta-5-3-ketosteroid isomerase from Pseudomonas testosteroni was determined by means of sedimentation equilibrium and exclusion chromatography over a wide range of enzyme concentrations in 0.2 M potassium phosphate buffer, pH 7.0. In addition, the sedimentation constant of the enzyme was determinded over an extended range of concentrations. The enzyme was found to have a molecular weight of 26,000 plus or equal to 1,000, suggesting that it is a dimer of identical or similar 13,400 molecular weight polypeptide chains. In the ultracentrifuge this dimeric species was found to undergo aggregation at enzyme concentrations above 2 mg per ml and dissociation at enzyme concentrations below 0.05 mg per ml. Exclusion chromatography studies indicate that under the conditions of chromatography the oligomeric enzyme is partially dissociated at enzyme concentrations in the range 0.2 to 0.002 mug per ml. These results suggest that under conditions of enzyme assay in 0.2 M potassium phosphate buffer, pH 7.0, isomerase is in a monomeric state of aggregation.  相似文献   

3.
Salt-dependent interconversion of inner histone oligomers.   总被引:1,自引:1,他引:0       下载免费PDF全文
The inner histone complex, extracted from chicken erythrocyte chromatin in 2 M NaCL AT pH 7.4, has been characterized by sedimentation equilibrium and sedimentation velocity. High speed sedimentation equilibrium studies indicate that in 2 M NaCl the inner histones are a weakly associating system with contributions from species ranging in molecular weight from dimer to octamer. The appearance of a single boundary (3.8S at 2 M NaCl) in sedimentation velocity studies conducted over a wide range of protein concentrations and ionic conditions indicates that the various histone oligomers present are in rapid equilibrium with one another. At higher salts the equilibrium is shifted to favor higher molecular weight species; in 4 M NaCl essentially all of the histone is octameric at protein concentrations above 0.2 mg/ml. The facile interconversion of histone oligomers suggests that small alterations in histone-histone interactions may be responsible for changes in nucleosome conformations during various biological processes.  相似文献   

4.
Abstract— The molecular weight of β nerve growth factor protein determined by sedimentation equilibrium in sodium acetate buffer, pH 40, and at protein concentrations around 0-5 mg/ml agrees with the value obtained from the amino acid sequence and confirms the dimeric character of the protein under these conditions. At pH values of 5.0 or greater, β nerve growth factor protein shows either partial dissociation into monomers or aggregation to higher polymers or both phenomena. The extent of dissociation or aggregation depends on buffer type and pH and is most pronounced at alkaline pH. The variation of molecular weight of β nerve growth factor with solvent conditions is similar to that of insulin or proinsulin. Removal of either the two COOH-terminal arginine residues or the two NH2-terminal octapeptide sequences from the protein has no effect on its solution properties at acid pH, the protein remaining a dimer. Species such as 2-5 S nerve growth factor or cyanogen bromide cleaved nerve growth factor which are partically deficient in COOH-terminal arginine residues and/or NH2-octapeptide or nonapeptide sequences are also dimers at pH40. The protein derivative which lacks the two NH2-terminal octapeptide sequence does not, like β-nerve growth factor, display dissociation or aggregation behavior at neutral pH, indicating that these sequences are involved in monomer-monomer interactions.  相似文献   

5.
Under lipid-free conditions, human apolipoprotein C-II (apoC-II) exists in an unfolded conformation that over several days forms amyloid ribbons. We examined the influence of the molecular chaperone, alpha-crystallin, on amyloid formation by apoC-II. Time-dependent changes in apoC-II turbidity (at 0.3 mg/ml) were suppressed potently by substoichiometric subunit concentrations of alpha-crystallin (1-10 microg/ml). alpha-Crystallin also inhibits time-dependent changes in the CD spectra, thioflavin T binding, and sedimentation coefficient of apoC-II. This contrasts with stoichiometric concentrations of alpha-crystallin required to suppress the amorphous aggregation of stressed proteins such as reduced alpha-lactalbumin. Two pieces of evidence suggest that alpha-crystallin directly interacts with amyloidogenic intermediates. First, sedimentation equilibrium and velocity experiments exclude high affinity interactions between alpha-crystallin and unstructured monomeric apoC-II. Second, the addition of alpha-crystallin does not lead to the accumulation of intermediate sized apoC-II species between monomer and large aggregates as indicated by gel filtration and sedimentation velocity experiments, suggesting that alpha-crystallin does not inhibit the relatively rapid fibril elongation upon nucleation. We propose that alpha-crystallin interacts stoichiometrically with partly structured amyloidogenic precursors, inhibiting amyloid formation at nucleation rather than the elongation phase. In doing so, alpha-crystallin forms transient complexes with apoC-II, in contrast to its chaperone behavior with stressed proteins.  相似文献   

6.
The sedimentation behavior of 125I-labeled gastrin has been studied as a function of Fe3+ ion concentration and pH. Both sedimentation velocity and sedimentation equilibrium experiments indicated that high-molecular-weight Fe3+-gastrin complexes were formed at pH 5.0 and pH 7.4. Self-association of gastrin alone was observed at pH values below 5.0. 125I-labeled gastrin bound to human serum apotransferrin at pH 7.4. Scatchard analysis of the gastrin-apotransferrin complex gave a Kd of approximately 6.4 microM at 37 degrees C, with two binding sites per molecule of apotransferrin. No significant binding of gastrin to diferric transferrin was observed under the same conditions. The binding of gastrin to apotransferrin was inhibited by NaCl. The results are consistent with the hypothesis that gastrin and transferrin act synergistically in the uptake of dietary iron by the gastrointestinal tract.  相似文献   

7.
1. The process of denaturation of the chicken muscle dimeric enzyme triosephosphate isomerase on addition of guanidinium chloride has been studied at pH 7.6, the pH at which the recovery of activity is optimal (100%) on removal of denaturant. Determinations of the sedimentation coefficient, intrinsic viscosity, molecular weight (by sedimentation equilibrium studies) and the absorption coefficient at 280 nm in various concentrations of guanidinium chloride concurred in showing a single, sharp transition at about 0.7 M guanidinium chloride at a protein concentration 1-5 mg/ml from the native enzyme to the dissociated, unfolded chains of the monomer. Relative fluorescent intensity measurements revealed a single transition at about 0.4 M guanidinium chloride at enzyme concentrations of about 0.05 mg/ml. 2. The process of denaturation in different guanidinium chloride concentrations was first order with respect to enzyme and about sixth order with respect to denaturant. 3. The rate of attainment of equilibrium during the renaturation obeyed second-order/first-order reversible kinetics. It was concluded that the rate-determining step in renaturation at pH 7.6 must be the association of two subunits.  相似文献   

8.
Dialysis-related amyloidosis is characterized by the deposition of insoluble fibrils of beta(2)-microglobulin (beta(2)-m) in the musculoskeletal system. Atomic force microscopy inspection of ex vivo amyloid material reveals the presence of bundles of fibrils often associated to collagen fibrils. Aggregation experiments were undertaken in vitro with the aim of reproducing the physiopathological fibrillation process. To this purpose, atomic force microscopy, fluorescence techniques, and NMR were employed. We found that in temperature and pH conditions similar to those occurring in periarticular tissues in the presence of flogistic processes, beta(2)-m fibrillogenesis takes place in the presence of fibrillar collagen, whereas no fibrils are obtained without collagen. Moreover, the morphology of beta(2)-m fibrils obtained in vitro in the presence of collagen is extremely similar to that observed in the ex vivo sample. This result indicates that collagen plays a crucial role in beta(2)-m amyloid deposition under physiopathological conditions and suggests an explanation for the strict specificity of dialysis-related amyloidosis for the tissues of the skeletal system. We hypothesize that positively charged regions along the collagen fiber could play a direct role in beta(2)-m fibrillogenesis. This hypothesis is sustained by aggregation experiments performed by replacing collagen with a poly-L-lysine-coated mica surface. As shown by NMR measurements, no similar process occurs when poly-L-lysine is dissolved in solution with beta(2)-m. Overall, the findings are consistent with the estimates resulting from a simplified collagen model whereby electrostatic effects can lead to high local concentrations of oppositely charged species, such as beta(2)-m, that decay on moving away from the fiber surface.  相似文献   

9.
The aggregation and dispersity of isolated bovine adrenal secretory vesicles (chromaffin granules) were studied by intensity fluctuation spectroscopy. The degree of dispersity and the Z-average translational diffusion coefficients were calculated from the autocorrelation functions of the intensity fluctuations in lase light scattered from the granules in solution. Granules purified by sedimentation through 0.3 M sucrose/Ficoll/2H2O showed greater dispersity than granules purified by sedimentation through 1.6 M sucrose. By monitoring the scattered light intensity and the diffusion coefficients of the granules, many of the difficulties encountered in the interpretation of absorbance measurements were avoided. Measurements over a range of granule concentrations in sucrose solutions (10 mM HEPES, pH 7.0), indicated that aggregation of the granules occurred at concentrations above 150 μg protein/ml. At low granule concentrations (15–30 μg protein/ml) Ca2+-induced aggregation was detected at a threshold of 2–10 mM calcium.  相似文献   

10.
It has been claimed that beta2-microglobulin (beta2-m) interacts with type I and type II collagen, and this property has been linked to the tissue specificity of the beta2-m amyloid deposits that target the osteo-articular system. The binding parameters of the interaction between collagen and beta2-m were determined by band shift electrophoresis and surface plasma resonance by using bovine collagen of type I and type II and various isoforms of beta2-m. Wild-type beta2-m binds collagen type I with a Kd of 4.1 x 10(-4) M and type II with 2.3 x 10(-3) M. By the BIAcore system we monitored the binding properties of the conformers of the slow phase of folding of beta2-m. The folding intermediates during the slow phase of folding do not display any significant difference with respect to the binding properties of the fully folded molecule. The affinity of beta2-m truncated at the third N-terminal residue does not differ from that reported for the wild-type protein. Increased affinity for collagen type I is found in the case of N-terminal truncated species lacking of six residues. The Kd of this species is 3.4 x 10 (-5) M at pH 7.4 and its affinity increases to 4.9 x 10(-6) M at pH 6.4. Fluctuations of the affinity caused by beta2-m truncation and pH change can cause modifications of protein concentration in the solvent that surrounds the collagen, and could contribute to generate locally a critical protein concentration able to prime the protein aggregation.  相似文献   

11.
Initiation factor MP was purified 1570-fold with 67% recovery of total activity present in 0.5 M KCl extracts of rabbit reticulocyte ribosomes. Initiation factor MP forms a ternary complex with Met-tRNAf and GTP or a binary complex with Met-tRNAf alone, the details of which are presented in the accompanying paper (Safer, B., Adams, S. L., Anderson. W. F., and Merrick, W. C. (1975) J. Biol. Chem. 250, 9076-9082). Initiation factor MP was homogeneous by the following criteria: (a) electrophoresis as a single band in gels of 5, 6, 7, 8, 9, and 10% acrylamide; (b) equilibration as a single band during isoelectric focusing; (c) sedimentation as a single symmetrical boundary during sedimentation velocity experiments; (d) linear plots of sedimentation equilibrium data; (e) symmetrical absorbance (at 280 nm) and activity profiles during DEAE-cellulose and Sephadex G-200 chromatography, and (f) symmetrical distribution of initiation factor MP during sucrose density gradient band sedimentation. The molecular weight of the initiation factor MP monomer (0.2 mg/ml) by low speed sedimentation equilibrium was 90,800. Calculations based on the Stokes radius and sedimentation velocity show the existence of relatively stable 90,000-dalton monomers or 180,000-dalton dimers at low (0.1 mg/ml) and high (9.75 mg/ml) concentrations of initiation factor MP, respectively. Electrophoresis in sodium dodecyl sulfate gels indicates that initiation factor MP monomer is composed of two noncovalently linked subunits with molecular weights of 52,000 and 34,000. Despite a relatively normal amino acid composition and an isoelectric point of 6.4, initiation factor MP behaves as a basic protein, eluting from phosphocellulose at 650 mM KCl (pH 7.9). Both ternary complex formation and methionyl-puromycin synthesis co-purify, indicating that a single protein is required for both activities.  相似文献   

12.
The influence of phosphate, ionic strength, temperature and enzyme concentration on the oligomeric structure of calf spleen purine nucleoside phosphorylase (PNP) in solution was studied by analytical ultracentrifugation methods. Sedimentation equilibrium analysis used to directly determine the enzyme molecular mass revealed a trimeric molecule with Mr = (90.6 +/- 2.1) kDa, regardless the conditions investigated: protein concentration in the range 0.02-1.0 mg/ml, presence of up to 100 mM phosphate and up to 200 mM NaCl, temperature in the range 4-25 degrees C. The sedimentation coefficient (6.04 +/- 0.02) S, together with the diffusion coefficient (6.15 +/- 0.11) 10(-7) cm2/s, both values obtained from the classic sedimentation velocity method at 1.0 mg/ml PNP concentration in 20 mM Hepes, pH 7.0, yielded a molecular mass of (90.2 +/- 1.6) kDa as expected for the trimeric enzyme molecule. Moreover, as shown by active enzyme sedimentation, calf spleen PNP remained trimeric even at low protein concentrations (1 microg/ml). Hence in solution, similar like in the crystalline state, calf spleen PNP is a homotrimer and previous suggestions for dissociation of this enzyme into more active monomers, upon dilution of the enzyme or addition of phosphate, are incorrect.  相似文献   

13.
The formation of deoxyhemoglobin was examined by measuring the heme spectral change that accompanies the aggregation of isolated alpha and beta chains. At low hemeconcentrations (less than 10(-5) M), tetramer formation can be described by two consecutive, second order reactions representing the aggregation of monomers followed by the association of alphabeta dimers. At neutral pH, the rates of monomer and dimer aggregation are roughly the same, approximately 5 X 10(5) M(-1) X(-1) at 20 degrees. Raising or lowering the pH results in a uniform decrease of both aggregation rates due presumably to repulsion of positively charged subunits at acid pH and repulsion of negatively charged subunits at alkaline pH. Addition of p-hydroxymercuribenzoate to alpha chains lowers the rate of monomer aggregation whereas addition of mercurials to the beta subunits appears to lower both the rate of monomer and the rate of dimer aggregation. At high heme concentrations (greater than 10(-5) M) or in the presence of organic phosphates, the rate of chain aggregation becomes limited, in part, by the slow dissociation of beta chain tetramers. In the case of inositol hexaphosphate, the rate of hemoglobin formation exhibits a bell-shaped dependence on phosphate concentration. When intermediate concentrations of inositol hexaphosphate (approximately 10(-4 M) are preincubated with beta subunits, a slow first order time course is observed and exhibits a half-time of about 8 min. As more inositol hexaphosphate is added, the chain aggregation reaction begins to occur more rapidly. Eventually at about 10(-2) M inositol hexaphospate, the time course becomes almost identical to that observed in the absence of phosphates. The increase in the velocity of the chain aggregation reaction at high phosphate concentrations suggests strongly that inositol hexaphosphate binds to beta monomers and, if added in sufficiently large amounts, promotes beta4 dissociation. A quantitative analysis of these results showed that the affinity of beta monomers for inositol hexaphosphate is the same as that of alphabeta dimers. Only when tetramers are formed, either alpha2beta2 or beta4, is a marked increase in affinity for inositol hexaphosphate observed.  相似文献   

14.
C J Westover  C L Stevens 《Biochemistry》1977,16(26):5819-5824
The lowest stages of polymerization of tobacco mosaic virus protein were studied by means of high-speed sedimentation equilibrium experiments. Several distinct modes of polymerization were found. At pH 7.1 the expected monomer-trimer-higher polymer equilibrium was observed--very little dimer was detected at this pH. At pH 7.5, however, a strong dimerization was observed--neither monomer nor trimer was detected at this pH. An octamer appeared to be the only species present other than the dimer. When 0.01 M beta-mercaptoethanol was added to the solvent pH 7.5, the dimer was dissociated, resulting in a monomer-trimer association. The dimerization may be the basis for the larger "doubled" polymers formed by the protein at alkaline pH, while the octamer may correspond to the 8S peak frequently observed in sedimentation velocity experiments at alkaline pH. On the other hand, the monomer-trimer-higher polymer equilibrium may correspond to the single helix formed by the protein at slightly acid pH and to the combination of 4S and 20S peaks seen in sedimentation velocity experiments at slightly acid pH.  相似文献   

15.
It has been confirmed by sedimentation equilibrium and sedimentation velocity experiments that des-(B26-B30)-insulin does not self-associate at neutral pH. Sedimentation equilibrium experiments at pH 7, 25 degrees C were conducted to investigate the effects of the structurally and physiologically important divalent cations Zn2+, Cd2+, Pb2+ and Ca2+ on the aggregation state of des-(B26-B30)-insulin (pig) in solution. It was found that all of these ions bring about association of this insulin analogue; Zn2+ and Cd2+ to a more marked degree than Pb2+ and Ca2+. The predominant species in solutions containing Zn2+ appear to be hexamers and hexameric aggregates, in those containing Cd2+, species up to and including tetramers, and in those containing Pb2+ and Ca2+, monomers and dimers of des-(B26-B30)-insulin appear to be the only species present. The possible significance of these findings, especially in relation to a role for Ca2+ in the action of insulin, is discussed.  相似文献   

16.
The dissociations of porcine heart mitochondrial, bovine heart mitochondrial, and porcine heart cytoplasmic malate dehydrogenase dimers (L-malate: NAD+oxidoreductase, EC 1.1.1.37) have been examined by Sephadex G-100 gel filtration chromatography and sedimentation velocity ultracentrifugation. The porcine mitochondrial enzyme was found to chromatograph as subunits when applied to a gel filtration column at a concentration of .02 muM or less at pH 7.0. The presence of coenzymes shifted the dissociation equilibrium at low enzyme concentrations in favor of dimer formation. Monomer formation was also favored when procine mitochondrial enzyme was incubated at pH 5.0 even at concentrations as high as 120 muM. This shift in equilibrium has been correlated with the increased rate and specificity of sulfhydryl residue modification with N-ethylmaleimide at pH 5.0 (Gregory, E.M., Yost, F.J.,Jr., Rohrbach, M.S., and Harrison, J.H. (1971)J. Biol. Chem. 246, 5491-5497). Bovine mitochondrial enzyme did not exhibit a concentration-dependent disociation under the conditions examined. However, at pH5.0 monomer formation was favored, and correlations could again be drawn with sulfhydryl residue modification (Gregory, E.M. (1975)J.Biol. Chem. 250, 5470-5474). In both mitochondrial enzymes, coenzyme binding was found capable of overcoming the effects of pH on the dissociation equilibrium, and dimer formation was favored. Unlike either of the above mentioned enzymes, porcine cytoplasmic malate dehydrogenase did not dissociate into its monomeric form under any conditions investigated.  相似文献   

17.
Connectin is an elastic protein of vertebrate striated muscle, and consists of doublet components, alpha and beta (also called titins 1 and 2). In the present study, beta-connectin isolated in the native state was investigated in order to characterize its molecular size and shape. The molecular weight was approximately 2.1 X 10(6) (SDS gel electrophoresis) or 2.7 X 10(6) (sedimentation equilibrium). The sedimentation coefficient (SO20, w) was 17S in 0.1 M phosphate buffer, pH 7.0. The intrinsic viscosity measured in an Ostwald-type viscometer was 1.8 dl/g. However, the viscosity was greatly dependent on the velocity gradient, and at a very low velocity gradient of 0.0007 s-1, a solution of connectin (0.3 mg/ml) showed a viscosity value of 17,000 cp. Flow birefringence measurements suggested a length distribution ranging from 300 to 450 nm. Electron microscopic observations revealed that connectin is a long flexible filament and the peaks of frequency of length distribution were at 150, 300, 450, and 600 nm. It was tentatively assumed that the connectin molecule is 300-400 nm long and 34-38 nm wide. It is likely that beta-connectin is derived from alpha-connectin, which has an apparent molecular weight of 2.8 X 10(6).  相似文献   

18.
Hydrodynamic studies of the solution properties of proteins and other biological macromolecules are often hard to interpret when the sample is present at a reasonably concentrated solution. The reason for this is that solutions exhibit deviations from ideal behaviour which is manifested as thermodynamic non-ideality. The range of concentrations at which this behaviour typically is exhibited is as low as 1–2 mg/ml, well within the range of concentrations used for their analysis by techniques such as small-angle scattering. Here we discuss thermodynamic non-ideality used previously used in the context of light scattering and sedimentation equilibrium analytical ultracentrifugation and apply it to the Guinier region of small-angle scattering data. The results show that there is a complementarity between the radially averaged structure factor derived from small-angle X-ray scattering/small-angle neutron scattering studies and the second virial coefficient derived from sedimentation equilibrium analytical ultracentrifugation experiments.  相似文献   

19.
The recently discovered cyclotides kalata B1 and kalata B2 are miniproteins containing a head-to-tail cyclized backbone and a cystine knot motif, in which disulfide bonds and the connecting backbone segments form a ring that is penetrated by the third disulfide bond. This arrangement renders the cyclotides extremely stable against thermal and enzymatic decay, making them a possible template onto which functionalities can be grafted. We have compared the hydrodynamic properties of two prototypic cyclotides, kalata B1 and kalata B2, using analytical ultracentrifugation techniques. Direct evidence for oligomerization of kalata B2 was shown by sedimentation velocity experiments in which a method for determining size distribution of polydisperse molecules in solution was employed. The shape of the oligomers appears to be spherical. Both sedimentation velocity and equilibrium experiments indicate that in phosphate buffer kalata B1 exists mainly as a monomer, even at millimolar concentrations. In contrast, at 1.6 mm, kalata B2 exists as an equilibrium mixture of monomer (30%), tetramer (42%), octamer (25%), and possibly a small proportion of higher oligomers. The results from the sedimentation equilibrium experiments show that this self-association is concentration dependent and reversible. We link our findings to the three-dimensional structures of both cyclotides, and propose two putative interaction interfaces on opposite sides of the kalata B2 molecule, one involving a hydrophobic interaction with the Phe6, and the second involving a charge-charge interaction with the Asp25 residue. An understanding of the factors affecting solution aggregation is of vital importance for future pharmaceutical application of these molecules.  相似文献   

20.
NAP-22, a myristoylated, anionic protein, is a major protein component of the detergent-insoluble fraction of neurons. After extraction from the membrane, it is readily soluble in water. NAP-22 will partition only into membranes with specific lipid compositions. The lipid specificity is not expected for a monomeric myristoylated protein. We have studied the self-association of NAP-22 in solution. Sedimentation velocity experiments indicated that the protein is largely associated. The low concentration limiting s value is approximately 1.3 S, indicating a highly asymmetric monomer. In contrast, a nonmyristoylated form of the protein shows no evidence of oligomerization by velocity sedimentation and has an s value corresponding to the smallest component of NAP-22, but without the presence of higher oligomers. Sedimentation equilibrium runs indicate that there is a rapidly reversible equilibrium between monomeric and oligomeric forms of the protein followed by a slower, more irreversible association into larger aggregates. In situ atomic force microscopy of the protein deposited on mica from freshly prepared dilute solution revealed dimers on the mica surface. The values of the association constants obtained from the sedimentation equilibrium data suggest that the weight concentration of the monomer exceeds that of the dimer below a total protein concentration of 0.04 mg/ml. Since the concentration of NAP-22 in the neurons of the developing brain is approximately 0.6 mg/ml, if the protein were in solution, it would be in oligomeric form and bind specifically to cholesterol-rich domains. We demonstrate, using fluorescence resonance energy transfer, that at low concentrations, NAP-22 labeled with Texas Red binds equally well to liposomes of phosphatidylcholine either with or without the addition of 40 mol% cholesterol. Thus, oligomerization of NAP-22 contributes to its lipid selectivity during membrane binding.  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