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.
Pre-steady-state studies of the isocitrate dehydrogenase reaction show that the rate constant for the hydride-transfer step is above 990s-1, and that both subunits of the enzyme are simulataneously active. After the fast formation of NADPH in amounts equivalent to the enzyme subunit concentration, the rate of NADPH formation is equal to the steady-state rate if the enzyme has been preincubated with isocitrate and Mg2+. If the enzyme has been preincubated with NADP+ and Mg2+, in 0.05 M-triethanolamine chloride buffer, pH 7.0, with the addition of 0.1 M-NaCl, the amount of NADPH formed in the fast phase is only 60% of the enzyme subunit concentration, and the turnover rate is at first lower than the steady-state rate. In 0.05 M-triethanolamine chloride buffer, pH 7.0, if the enzyme is preincubated with NADP+ or NADPH, the turnover rate increases 3-fold to reach the steady-state rate after about 5 s. Preincubation of the enzyme with isocitrate and Mg2+ abolishes this lag phase, the steady-state rate being reached at once. It is suggested that the enzyme exists in at least two conformational forms with different activities, and that the lag phase represents the transition (k = 0.4s-1) from a form with low activity to the fully active enzyme, induced by the binding of isocitrate and Mg2+.  相似文献   

2.
Glutathione reductase from S. cerevisiae (EC 1.6.4.2) catalyzes the NADPH oxidation by glutathione in accordance with a "ping-pong" scheme. The catalytic constant kcat) is 240 s-1 (pH 7.0, 25 degrees C); kcat for the diaphorase reaction is 4-5 s-1. The enzyme activity does not change markedly at pH 5.5-8.0. At pH less than or equal to 7.0, NADP+ acts as a competitive inhibitor towards NADPH and as a noncompetitive inhibitor towards glutathione. NADP+ increases the diaphorase activity of the enzyme. The maximal activity is observed, when the NADP+/NADPH ratio exceeds 100. At pH 8.0, NADP+ acts as a mixed type inhibitor during the reduction of glutathione. High concentrations of NADP+ also inhibit the diaphorase activity due to the reoxidation of the reduced enzyme by NADP+ at pH 8.0. The redox potential of glutathione reductase calculated from the inhibition data is--306 mV (pH 8.0). Glutathione reductase reduces quinoidal compounds in an one-electron way. The hyperbolic dependence of the logarithm of the oxidation constant on the one electron reduction potential of quinone is observed. It is assumed that quinones oxidize the equilibtium fraction of the two-electron reduced enzyme containing reduced FAD.  相似文献   

3.
Sheep liver 6-phosphogluconate dehydrogenase is shown to be inactivated by diethylpyrocarbonate in a biphasic manner at pH 6.0, 25 degrees C. After allowing for the hydrolysis of the reagent, rate constants of 56 M-1 s-1 and 11.0 M-1 s-1 were estimated for the two processes. The complete reactivation of partially inactivated enzyme by neutral hydroxylamine, the elimination of the possibility that modification of cysteine or tyrosine residues are responsible for inactivation, and the magnitudes of the rate constants for inactivation relative to the experimentally determined value for the reaction of diethylpyrocarbonate with N alpha-acetylhistidine (2.2 M-1 s-1), all suggested that enzyme inactivation occurs solely by modification of histidine residues. Comparison of the experimental plot of residual fractional activity versus the number of modified histidine residues per subunit with simulated plots for three hypothetical models, each predicting biphasic kinetics, indicated that inactivation results from the modification of at most one essential histidine residue per subunit, although it appears that other (non-essential) histidines react independently. This histidine is thought to be His-242 and is present in the active site. Evidence in support of its role in catalysis is briefly discussed. Both 6-phosphogluconate and organic phosphate protect against inactivation, and a kinetic analysis of the protection indicated a dissociation constant of 2.1 X 10(-6) M for the enzyme--6-phosphogluconate complex. NADP+ also protected, but this might be due, at least in part, to a reduction in the effective concentration of diethylpyrocarbonate.  相似文献   

4.
6-Phosphogluconate dehydrogenase is the pivotal enzyme that links the gluconate route and the oxidative phase of the pentose phosphate pathway in Schizosaccharomyces pombe. The enzyme differs from the known 6-phosphogluconate dehydrogenases of other sources in that the Schizosaccharomyces enzyme is tetrameric having a subunit mass of 38 kDa, that it requires NADP+ obligatorily for activity, and that it can be activated by divalent metal ions such as Co2+ and Mn2+. Steady-state kinetic studies were undertaken. Initial rate and product inhibition results suggest that 6-phosphogluconate dehydrogenase from Schizosaccharomyces pombe catalyzes NADP(+)-linked oxidative decarboxylation of 6-phosphogluconate by an equilibrium random mechanism with two independent binding sites, namely one site for the nicotinamide coenzyme, NADP+/NADPH, and another site for 6-phosphogluconate-D-ribulose-5-phosphate and for CO2. Studies of pH dependence implicated a basic residue with a pK value of 7.4 in the binding of 6-phosphogluconate and an acidic residue with a pK value of 6.7 in the cation-mediated interaction of NADP+ with the enzyme.  相似文献   

5.
The kinetics of a single turnover of enzyme-catalysed oxidative decarboxylation have been studied by mixing a stoichiometric complex of enzyme, isocitrate and Mg2+ with large concentrations of NADP+ in a stopped-flow apparatus, and monitoring the formation of NADPH by fluorescence measurements. A transient is revealed that exhibits enhanced nucleotide fluorescence and is not detectable by light absorption measurements. The results obtained with the largest NADP+ concentrations, such that the product NADPH is largely displaced from its enzyme complex, show that a step that precedes the release of free NADPH is rate-limiting in the oxidative decarboxylation reaction under conditions of catalytic cycling. The rate constants for this step, tentatively identified as the formation of the complex of enzyme, Mg2+ and NADPH from a precursor NADPH-containing complex, and for the dissociation of NADPH from this complex have been estimated from the integrated rate equation for a simple model for the product phase of the reaction, by methods of nonlinear regression analysis. In line with the conclusions from the preceding paper, it is suggested that formation of an abortive complex of enzyme, Mg2+, isocitrate and NADPH under catalytic cycling conditions serves to by-pass the potentially rate-limiting dissociation of NADPH from the enzyme-Mg2+-NADPH complex.  相似文献   

6.
The fluorescence quantum yield of NADPH is enhanced in its complex with 6-phospho-gluconate dehydrogenase, and a further enhancement in the presence of excess 6-phospho-gluconate shows that an abortive ternary complex is formed. There is marked energy transfer from aromatic residues in the enzyme to NADPH in the complexes, as indicated by an excitation maximum at 280 nm in the fluorescence excitation spectrum of the complex. The coenzyme fluorescence enhancement has been used to determine the dissociation constant for NADPH in the binary and ternary complexes, and the stoichiometry of the complexes, from the results of fluorescence titrations. A new method of analysis of fluorescence titration data is described. The results show that each subunit of the dimeric enzyme binds NADPH independently and with the same affinity. The dissociation constant for the enzyme-coenzyme complex, in phosphate buffer, pH 7.0, is 5.7 μm; the dissociation constant for NADPH in the ternary complex with 6-phosphogluconate is 7.0 μm.  相似文献   

7.
A method is described for the isolation and purification of 6-phosphogluconate dehydrogenase from pig liver. The molecular weight is estimated at 83,000 and that of the subunits is 42,000 as determined by gel electrophoresis. The pH maximum is 8.5 in 50 mM glycine/NaOH buffer and from 7.5 to 10 in 50 mM phosphate buffer at 30 degrees. Magnesium ion is not required for activity and acts as an inhibitor at concentrations above 20 mM. A cellular fractionation study indicates that this enzyme is located almost entirely within the soluble portion of the cytoplasm. Kinetic studies have been done in 50 mM glycine buffer, pH 8.5, at 30 degrees. The data are consistent with a sequential mechanism in which NADP+ is added first, followed by 6-phosphogluconate, and the products are released in the order, CO2, ribulose 5-phosphate, and NADPH. The Michaelis constant is 13.5 muM for 6-phosphogluconate. Dissociation constants are 4.8 muM for NADP+ and 5.1 muM for NADPH.  相似文献   

8.
Steady-state kinetic studies including initial velocity, NADPH product inhibition, dead-end inhibition, and combined dead-end and product inhibition measurements with purified rat liver glucose-6-phosphate dehydrogenase indicate a sequential and obligatory addition of substrates in the order of NADP+, glucose-6-P for the catalytic pathway at pH 8.0. Although instability of 6-phosphoglucono-delta-lactone precluded product inhibition experiments which might directly exclude an enzyme-6-phosphoglucono-delta-lactone complex, the absence of an enzyme-glucose-6-P complex suggests that the enzyme-lactone product is unlikely and the release of products is also ordered, with NADPH released last. Consideration of the kinetic constants (Ka = 2.0 muM, Kiq = 13 muM) and cellular concentration of the substrates and products suggests extensive inhibition of the enzyme in vivo and control by the NADPH/NADP+ ratios. Circular dichroism spectra of the enzyme in 20 mM phosphate buffer at pH 7.0 and 25 degrees C indicate 51% helix and 33% pleated sheet structures which is considerably different from results (14% helix) with yeast enzymes.  相似文献   

9.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  相似文献   

10.
The pigeon liver fatty acid synthetase complex (14 S) is dissociated in low ionic strength buffer containing dithiothreitol to form a half-molecular weight subunits (9 S) which are completely inactive for the synthesis of saturated fatty acids. The dithiothreitol-protected (reduced) subunits are rapidly reassociated and reactivated to form the active enzyme complex, not only by an increase in salt concentration but also by micromolar concentrations of NADP+ or NADPH. Increases in KCl or NADPH concentration result in an increase in the extent of reactivation (equilibrium) with no change in the over-all rate of the reaction or the half-life ofreactivation of the enzyme. The extent (equilibrium) of reactivation of the enzyme is the same in 0.2 M potassium phosphate buffer, pH 7.0; 0.2 M KCl in 5 mM Tris-35 mM glycine buffer, PH 8.3; and 50 muM NADP+ or NADPH in the Tris-glycine buffer. The extent and rate of reactivation of the enzyme is dependent not only on ionic strength and NADPH concentration, but also on pH and temperature. Reactivation with 0.2 M KCl is optimal between pH 7.3 and 8.5. At higher and lower pH values the rate and extent of reactivation are lowered. The rate and extent of reactivation are also decreased as the temperature is lowered below 10 degrees. At 0 degrees there is little reactivation of enzyme activity. However, in the presence of 0.2 M KCl containing 15 to 40% glycerol at 0 degrees, reactivation of the enzyme is about 50% complete. The rate of reactivation of enzyme in the presence of KCl or NADPH conforms to first order kinetics. This result suggests that the subunits first combine to form an inactive complex which is subsequently transformed to an enzymatically active complex. Evidence for the presence of inactive complex was obtained in experiments carried out in 0.2 M KCl at pH 6.0, and in 0.2 M KCl at pH 8.3, at both 6 and 3 degrees. Under these conditions the amount of complex observed upon ultracentrifugation was greater than expected from determinations of enzyme activity. The above findings suggest that ionic and hydrophobic interactions, and possibly the water structure surrounding the interacting sites, are of prime importance in reassociation and reactivation of enzyme. In addition, NADP+ and NADPH have very specific effects in bringing about reassociation and in maintaining the structural integrity of the multienzyme complex.  相似文献   

11.
The activity of glucose-6-phosphate dehydrogenase (EC 1.1.1.49) FROM SPINACH CHLOROPLASTS IS STRONGLY REGULATED BY THE RATIO OF NADPH/NADP+, with the extent of this regulation controlled by the concentration of ribulose 1,5-diphosphate. Other metabolites of the reductive pentose phosphate cycle are far less effective in mediating the regulation of the enzyme activity by NADPH/NADP+ ratio. With a ratio of NADPH/NADP+ of 2, and a concentration of ribulose 1,5-diphosphate of 0.6 mM, the activity of the enzyme is completely inhibited. This level of ribulose 1,5-diphosphate is well within the concentration range which has been reported for unicellular green algae photosynthesizing in vivo. Ratios of NADPH/NADP+ of 2.0 have been measured for isolated spinach chloroplasts in the light and under physiological conditions. Since ribulose 1,5-diphosphate is a metabolite unique to the reductive pentose phosphate cycle and inhibits glucose-6-phosphate dehydrogenase in the presence of NADPH/NADP+ ratios found in chloroplasts in the light, it is proposed that regulation of the oxidative pentose phosphate cycle is accomplished in vivo by the levels of ribulose 1,5-diphosphate, NADPH, and NADP+. It already has been shown that several key reactions of the reductive pentose phosphate cycle in chloroplasts are regulated by levels of NADPH/NADP+ or other electron-carrying cofactors, and at least one key-regulated step, the carboxylation reaction is strongly affected by 6-phosphogluconate, the metabolic unique to the oxidative pentose phosphate cycle. Thus there is an interesting inverse regulation system in chloroplasts, in which reduced/oxidized coenzymes provide a general regulatory mechanism. The reductive cycle is activated at high NADPH/NADP+ ratios where the oxidative cycle is inhibited, and ribulose 1,5-diphosphate and 6-phosphogluconate provide further control of the cycles, each regulating the cycle in which it is not a metabolite.  相似文献   

12.
A kinetic scheme is presented for Lactobacillus casei dihydrofolate reductase that predicts steady-state kinetic parameters. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. Two major features of this kinetic scheme are the following: (i) product dissociation is the rate-limiting step for steady-state turnover at low pH and follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex; (ii) the rate constant for hydride transfer from NADPH to dihydrofolate (H2F) is rapid (khyd = 430 s-1), favorable (Keq = 290), and pH dependent (pKa = 6.0), reflecting ionization of a single group. Not only is this scheme identical in form with the Escherichia coli kinetic scheme [Fierke et al. (1987) Biochemistry 26, 4085] but moreover none of the rate constants vary by more than 40-fold despite there being less than 30% amino acid homology between the two enzymes. This similarity is consistent with their overall structural congruence. The role of Trp-21 of L. casei dihydrofolate reductase in binding and catalysis was probed by amino acid substitution. Trp-21, a strictly conserved residue near both the folate and coenzyme binding sites, was replaced by leucine. Two major effects of this substitution are on (i) the rate constant for hydride transfer which decreases 100-fold, becoming the rate-limiting step in steady-state turnover, and (ii) the affinities for NADPH and NADP+ which decrease by approximately 3.5 and approximately 0.5 kcal mol-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The reactions of NADPH oxidation by quinones and inorganic complexes catalyzed by NADPH: adrenodoxin reductase were studied. The catalytic constant for the enzyme at pH 7.0 is 20-25 s-1; the oxidative constants for the quinones vary from 5 X 10(5) to 1.1 X 10(3) M-1 s-1 and show an increase with a rise in the one-electron acceptor reduction potential. The mode of adrenodoxin reductase interaction with oxyquinones differs from that of the enzyme interaction with alkyl-substituted quinones and inorganic complexes. NADPH competitively inhibits electron acceptors, whereas NADP+ is a competitive inhibitor of NADPH and a uncompetitive inhibitor of electron acceptors. (Ki = 25 microM). The depth of FAD incorporation into the enzyme molecule as calculated according to the outer sphere electron transfer theory is 6.1 A.  相似文献   

14.
Adrenodoxin reductase (EC 1.18.1.2) catalyzes the oxidation of NADPH by 1.4-benzoquinone. The catalytic constant of this reaction at pH 7.0 is equal to 25-28 s-1. NADP+ acts as the mixed-type nonlinear inhibitor of enzyme increasing Km of NADPH and decreasing catalytic constant. NADP+ and NADPH act as mutually exclusive inhibitors relative to reduced adrenodoxin reductase. The patterns of 2',5'-ADP inhibition are analogous to that of NADP+. These data support the conclusion about the existence of second nicotinamide coenzyme binding centre in adrenodoxin reductase.  相似文献   

15.
S A Adediran 《Biochimie》1991,73(9):1211-1218
The steady-state kinetics of normal human erythrocyte glucose-6-phosphate dehydrogenase (D-glucose-6-phosphate: NADP+ oxidoreductase, EC 1.1.1.49) dimers were studied as a function of pH and temperature. Inhibition studies using glucosamine 6-phosphate, NADPH and p-hydroxymercuribenzoate (P-OHMB) were also carried out at pH 8.0. The existence of two binding sites on the enzyme with a transition from low to high affinity for NADP+ when NADP+ concentration is increased is indicated by the nonlinear Lineweaver-Burk plots and sigmoid kinetic patterns. NADPH inhibition was found to be competitive with respect to NADP+ and non-competitive with respect to glucose-6-phosphate. Logarithmic plot of Vmax against pH and inactivation by P-OHMB indicate the participation in the reaction mechanism of imidazolium group of histidine and sulhydryl groups. The initial velocity and product inhibition data gave results which are consistent with the dimeric enzyme following an ordered sequential mechanism. A possible random mechanism is ruled out by the inhibition results of glucosamine 6-phosphate.  相似文献   

16.
Association and dissociation rate constants obtained by stopped-flow spectroscopy have permitted definition of a kinetic scheme for recombinant human dihydrofolate reductase that correctly predicts full time course kinetics of the enzymatic reaction over a wide range of substrate and product concentrations. The scheme is complex compared with that for the bacterial enzyme and involves branched pathways. It successfully accounts for observed rapid hysteresis preceding steady state and for the nonhyperbolic dependence of steady-state rate on substrate and product concentrations. The major branch point in the catalytic cycle occurs at E.NADP.H4folate because either NADP or H4folate can dissociate from the ternary product complex (koff = 84 s-1 and 46 s-1, respectively). The rate of conversion of enzyme-bound substrates to products is very fast (k = 1360 s-1) and nearly unidirectional (Kequ = 37) so that other steps limit the catalytic rate. At saturating substrate concentrations these steps include release of NADP and H4folate from E.NADP.H4folate and release of products from the two abortive complexes E.NADPH.H4folate (koff = 225 s-1) and E.NADP.H4folate (koff = 4.6 s-1). Since NADP dissociates slowly from E.NADP.H2folate nearly 90% of the enzyme accumulates as this complex at steady state. Nonetheless, the catalytic rate is maintained at 12 s-1 by rapid flux of a small portion of the enzyme through an alternate branch. At physiological concentrations of substrates and products the steady-state rate is limited primarily by the rate of H2folate binding to E.NADPH so that the enzyme is extremely efficient.  相似文献   

17.
Streptococcus faecalis grown with glucose as the primary energy source contains a single, nicotinamide adenine dinucleotide phosphate (NADP)-specific 6-phosphogluconate dehydrogenase. Extracts of gluconate-adapted cells, however, exhibited 6-phosphogluconate dehydrogenase activity with either NADP or nicotinamide adenine dinucleotide (NAD). This was shown to be due to the presence of separate enzymes in gluconate-adapted cells. Although both enzymes catalyzed the oxidative decarboxylation of 6-phosphogluconate, they differed from one another with respect to their coenzyme specificity, molecular weight, pH optimum, K(m) values for substrate and coenzyme, and electrophoretic mobility in starch gels. The two enzymes also differed in their response to certain effector ligands. The NADP-linked enzyme was specifically inhibited by fructose-1,6-diphosphate, but was insensitive to adenosine triphosphate (ATP) and certain other nucleotides. The NAD-specific enzyme, in contrast, was insensitive to fructose-1,6-diphosphate, but was inhibited by ATP. The available data suggest that the NAD enzyme is involved primarily in the catabolism of gluconate, whereas the NADP enzyme appears to function in the production of reducing equivalents (NADPH) for use in various reductive biosynthetic reactions.  相似文献   

18.
Structure and properties of malic enzyme from Bacillus stearothermophilus   总被引:3,自引:0,他引:3  
The malic enzyme (EC 1.1.1.38) gene of Bacillus stearothermophilus was cloned in Escherichia coli, and the enzyme was purified to homogeneity from the E. coli clone. In addition to the NAD(P)-dependent oxidative decarboxylation of L-malate, the enzyme catalyzes the decarboxylation of oxalacetate. The enzyme is a tetramer of Mr 200,000 consisting of four identical subunits of Mr 50,000. The pH optima for malate oxidation and pyruvate reduction are 8.0 and 6.0, respectively; and the optimum temperature is 55 degrees C. The enzyme strictly requires divalent metal cations for its activity, and the activity is enhanced 5-7 times by NH4+ and K+. Kinetic study shows that the values of the dissociation constant of the enzyme-coenzyme complex are 77 microM for NAD and 1.0 mM for NADP, indicating that the enzyme has a higher affinity for NAD than for NADP. The nucleotide sequence of the gene and its flanking regions was also found. A single open reading frame of 1434 base pairs encoding 478 amino acids was concluded to be that for the malic enzyme gene because the amino acid composition of the enzyme and the sequence of 16 amino acids from the amino terminus of the enzyme agreed well with those deduced from this open reading frame.  相似文献   

19.
The mitochondrial NADP-dependent malic enzyme (EC 1.1.1.40) was purified about 300-fold from cod Gadus morhua heart to a specific activity of 48 units (mumol/min)/mg at 30 degrees C. The possibility of the reductive carboxylation of pyruvate to malate was studied by determination of the respective enzyme properties. The reverse reaction was found to proceed at about five times the velocity of the forward rate at a pH 6.5. The Km values determined at pH 7.0 for pyruvate, NADPH and bicarbonate in the carboxylation reaction were 4.1 mM, 15 microM and 13.5 mM, respectively. The Km values for malate, NADP and Mn2+ in the decarboxylation reaction were 0.1 mM, 25 microM and 5 microM, respectively. The enzyme showed substrate inhibition at high malate concentrations for the oxidative decarboxylation reaction at pH 7.0. Malate inhibition suggests a possible modulation of cod heart mitochondrial NADP-malic enzyme by its own substrate. High NADP-dependent malic enzyme activity found in mitochondria from cod heart supports the possibility of malate formation under conditions facilitating carboxylation of pyruvate.  相似文献   

20.
A kinetic scheme is presented for Escherichia coli dihydrofolate reductase that predicts steady-state kinetic parameters and full time course kinetics under a variety of substrate concentrations and pHs. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. The binding kinetics suggest that during steady-state turnover product dissociation follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex. This step, H4F dissociation from the E X NADPH X H4F ternary complex, is proposed to be the rate-limiting step for steady-state turnover at low pH because koff = VM. The rate constant for hydride transfer from NADPH to dihydrofolate (H2F), measured by pre-steady-state transients, has a deuterium isotope effect of 3 and is rapid, khyd = 950 s-1, essentially irreversible, Keq = 1700, and pH dependent, pKa = 6.5, reflecting ionization of a single group in the active site. This scheme accounts for the apparent pKa = 8.4 observed in the steady state as due to a change in the rate-determining step from product release at low pH to hydride transfer above pH 8.4. This kinetic scheme is a necessary background to analyze the effects of single amino acid substitutions on individual rate constants.  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