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 437 毫秒
1.
A graphical method for analysing enzyme data to obtain kinetic parameters, to identify the types of inhibition and the enzyme mechanisms is described. The method consists of plotting experimental data as v/(V(0)-v) versus 1/(I) at different substrate concentrations. I is the inhibitor concentration; V(0) and v are the initial rates of enzyme reaction attained by the system in the presence of a fixed amount of substrate and in the absence and presence of inhibitor respectively. Complete inhibition gives straight lines that pass through the origin while partial inhibition gives straight lines that converge on the 1/I-axis at a point away from the origin. With uncompetitive inhibition the slopes of the lines decrease with increasing substrate concentration. The kinetic parameters K(m), K'(i) and beta (degree of partiality) can best be determined from respective secondary plots of slope and intercept versus reciprocal of substrate concentration.  相似文献   

2.
A graphical method for analyzing enzyme data to obtain kinetic parameters, and to identify the types of inhibition and the enzyme mechanisms, is described. The method consists of plotting experimental data as nu/(V0 - nu) vs 1/(I) at different substrate concentrations. I is the inhibitor concentration; V0 and nu are the rates of enzyme reaction attained by the system in the presence of a fixed amount of substrate, and in the absence and presence of inhibitor, respectively. Complete inhibition gives straight lines that go through the origin; partial inhibition gives straight lines that converge on the 1-I axis, at a point away from the origin. For competitive inhibition, the slopes of the lines increase with increasing-substrate concentration; with noncompetitive inhibition, the slopes are independent of substrate concentration; with uncompetitive inhibition, the slopes of the lines decrease with increasing substrate concentrations. The kinetic parameters, Km, Ki, Ki', and beta (degree of partiality) can best be determined from respective secondary plots of slope and intercept vs substrate concentration, for competitive and noncompetitive inhibition mechanism or slope and intercept vs reciprocal substrate concentration for uncompetitive inhibition mechanism. Functional consequencs of these analyses are represented in terms of specific enzyme-inhibitor systems.  相似文献   

3.
A new simple graphical method is described for the determination of inhibition type and inhibition constants of an enzyme reaction without any replot. The method consists of plotting experimental data as (Vv)/v versus the inhibitor concentration at two or more concentrations of substrate, where V and v represent the maximal velocity and the velocity in the absence and presence of inhibitor with given concentrations of the substrate, respectively. Competitive inhibition gives straight lines that converge on the abscissa at a point where [I]?=??Ki. Uncompetitive inhibition gives parallel lines with the slope of 1/K’i. For mixed type inhibition, the intersection in the plot is given by [I]?=??Ki and (V–v)/v?=??Ki/K’i in the third quadrant, and in the special case where Ki?=?K’i (noncompetitive inhibition) the intersections occur at the point where [I]?=??Ki and (Vv)/v?=??1. The present method, the “quotient velocity plot,” provides a simple way of determining the inhibition constants of all types of inhibitors.  相似文献   

4.
gamma-glutamyl Transferase fron Sheep brain cortex capillaries was studied from the point of view of transport of aminoacids across blood brain barrier. Excess substrate inhibition was competitive and observed both with donor (glutathione) and various acceptors (methionine, alanine, tryptophan) but not with arginine. Excess glutathione inhibition of transfer reaction is concomitant with an increase of total reaction (transfer + hydrolysis + autotranspeptidation). With regard to aminoacids, the greater the K'm the stronger the inhibition. This inhibition is the result of formation of a dead complex. Lineweaver-Burk plots 1/v versus 1/[acceptor] give straight lines meeting at the same point, whereas 1/v verus 1/[donor] plots are roughly parallel for high aminoacid concentrations and become secant for the low ones. Replots of slopes vs. 1/[acceptor] are not linear: the lower the aminoacid affinity the more pronounced the slope replot curvature. Thus kinetic patterns are consistent with a branched ping-pong mechanism including a ternary complex (Enzyme-acceptor-H2O) at high or low relative concentration, which balances the two branches. The estimated value of kinetic parameters does not support the hypothesis of major implication of the enzyme in brain uptake of aminoacids.  相似文献   

5.
P F Fitzpatrick 《Biochemistry》1991,30(15):3658-3662
The steady-state kinetic mechanism for rat tyrosine hydroxylase has been determined by using recombinant enzyme expressed in insect tissue culture cells. Variation of any two of the three substrates, tyrosine, 6-methyltetrahydropterin, and oxygen, together at nonsaturating concentrations of the third gives a pattern of intersecting lines in a double-reciprocal plot. Varying tyrosine and oxygen together results in a rapid equilibrium pattern, while the other substrate pairs both fit a sequential mechanism. When tyrosine and 6-methyltetrahydropterin are varied at a fixed ratio at different oxygen concentrations, the intercept replot is linear and the slope replot is nonlinear with a zero intercept, consistent with rapid equilibrium binding of oxygen. All the replots when oxygen is varied in a fixed ratio with either tyrosine or 6-methyltetrahydropterin are nonlinear with finite intercepts. 6-Methyl-7,8-dihydropterin and norepinephrine are competitive inhibitors versus 6-methyltetrahydropterin and noncompetitive inhibitors versus tyrosine. 3-Iodotyrosine, a competitive inhibitor versus tyrosine, shows uncompetitive inhibition versus 6-methyltetrahydropterin. At high concentrations, tyrosine is a competitive inhibitor versus 6-methyltetrahydropterin. These results are consistent with an ordered kinetic mechanism with the order of binding being 6-methyltetrahydropterin, oxygen, and tyrosine and with formation of a dead-end enzyme-tyrosine complex. There is no significant primary kinetic isotope effect on the V/K values or on the Vmax value with [3,5-2H2]tyrosine as substrate. No burst of dihydroxyphenylalanine production is seen during the first turnover. These results rule out product release and carbon-hydrogen bond cleavage as rate-limiting steps.  相似文献   

6.
In a model system derived from Escherichia coli, Ac[3H]Phe-puromycin is produced in a pseudo-first-order reaction between the preformed Ac[3H]Phe-tRNA-poly(U)-ribosome complex (complex C) and excess puromycin [Kalpaxis et al. Eur. J. Biochem. 154, 267, 1986]. Amicetin and gougerotin inhibit this reaction to various degrees depending on whether or not complex C is allowed to interact with the inhibitor (I) prior to the addition of puromycin (S). The kinetic analysis shows a phase where competitive inhibition can be observed provided that S and I are added simultaneously. After preincubating C with I, the inhibition becomes of the mixed non-competitive type. The Ki (the dissociation constant of the CI complex), calculated from the competitive plot, is 20.0 microM for amicetin and 15.0 microM for gougerotin. This inhibition constant (Ki) cannot distinguish amicetin from gougerotin. Its acceptance as a criterion of potency does not explain why after preincubation amicetin proves to be a stronger inhibitor than gougerotin. The determination of the apparent catalytic rate constants of peptidyltransferase at various inhibitor concentrations and the appropriate replotting of these rate constants distinguish amicetin from gougerotin. A new approach for evaluating the potency of these inhibitors is proposed. The familiar Ki is supplemented with an apparent kinetic constant obtained from a replot in which the intercepts of the double-reciprocal plots (1/kobs versus 1/[S]) are plotted versus the inhibitor concentration.  相似文献   

7.
An equation is found relating the fractional activity, (v/v0), of an enzyme assay mixture to the total concentrations of metalloenzyme, active site metal ion, metal-binding ligand and substrate and the stability constants of the complexes present. When (v/v0) is measured as a function of the total ligand concentration, this equation offers a way of data-plotting which yields straight lines and permits the calculation of the metal-binding constant KME from either the slope or the intercept, provided that mixed complexes (enzyme-metal ion-ligand) do not contribute significantly to the change in (v/v0). Since deviations from linearity occur in the latter case, the proposed inhibition plot serves as a diagnostic tool for the recognition of such complexes. Application to the inhibition of thermolysin by 1,10-phenanthroline gives a value of 2.1 × 1011m−1 for KZnE, the binding constant of the active site zinc ion, at pH 7.50, 25°C and ionic strength 0.1. The equation also allows the rapid calculation of the ligand concentration necessary to attain a desired degree of inhibition when the total enzyme and active site metal ion concentrations of the solution are known.  相似文献   

8.
We have been successful in building a mathematical model that fits both the germination rate and the total number of seeds that germinate as a function of time. This mathematical model is the same autocatalytic reaction model that describes biochemical reactions in which enzymes play an important role. The model gives values for the initial concentration of two enzymes. From these initial enzyme concentrations an equilibrium constant is calculated and the thermodynamic model gives the change in enthalpy, entropy, free energy and the activation energy. A plot of the natural logarithm of the equilibrium constant as a function of the reciprocal of the absolute temperature gives two straight lines. The change of enthalpy for the process below 33 °C differs considerably to the change above 33 °C. The free energy as a function of the absolute temperature gives a straight line from which the change in entropy is calculated. The activation energy is determined from the slope of the natural logarithm of the rate constant as a function of the reciprocal of the absolute temperature.  相似文献   

9.
A new procedure to characterize reversible dead-end inhibitors is presented. Preliminary identification of the inhibitor type is made by plotting vo/vi against the inhibitor concentration at different substrate concentrations. The inhibition constants for competitive, uncompetitive and mixed dead-end inhibitors are determined by secondary plots of l/(slope) vs [S], l/(slope) vs l/[S] and (slope)(Ks + [S] vs [S] respectively. These secondary plots render straight lines only for their corresponding type of inhibitor. For noncompetitive inhibitors all the secondary plots used yield straight lines. Therefore, the application of this plotting procedure leads to unambiguous diagnosis of the inhibitor type. An important feature of the procedure presented here is that the variable used (vo/vi) is independent on Vmax values. Therefore, experimental values obtained from enzyme preparations showing significant differences in their specific activities -i.e. enzyme coming from different purification steps- can be used.  相似文献   

10.
1. A theoretical appraisal of the alternative pathway mechanism for a two-substrate enzyme shows that this mechanism is capable of giving rise to apparent substrate inhibition or substrate activation (Dalziel, 1958). It has now been shown that these phenomena may occur simultaneously in the following ways. With certain relationships between the kinetic parameters and the constant concentration of one substrate, A, the plot of initial rate, v, against the concentration of the other substrate, B, may show substrate ;activation' at low concentrations of B and substrate ;inhibition' at high concentrations of B. In other circumstances the plot of v against [B], with [A] constant, may be sigmoid (substrate activation), whereas the plot of v against [A], with [B] constant, may pass through a maximum (substrate inhibition). 2. Kinetic data for phosphofructokinase are of the latter type and it is suggested that the mechanism of this enzyme may involve a kinetically preferred pathway. It is emphasized that the phenomena of substrate inhibition and activation need not necessarily involve more than one binding site for each substrate on the enzyme molecule, nor more than one monomer per molecule.  相似文献   

11.
The method of kinetic analysis is developed to obtain the maximum velocity (Vm), the Michaelis constant (Km) and the parameters characterizing the inhibitors in an impure enzyme reaction, contaminated with one of four types of inhibitor (competitive, noncompetitive, uncompetitive and mixed-type). Although the reaction rate decreases with the increasing concentration of the enzyme sample containing an inhibitor, the double-reciprocal plot of the rate against the sample concentration becomes linear. The slopes of these linear plots at several different concentrations of substrate provide Km and the specific enzyme activity, which is proportional to Vm, in the sample. These linear straight lines intersect in a point, of which the coordinates give the unique parameters for the inhibitor. To prove the validity of this kinetic method, the model experiments were carried out with acetylcholinesterase and its inhibitors, phenyltrimethylammonium and trimethylammonium. The present method was applied to the measurement of the specific activity of galactosylceramide galactosidase in the mouse cerebral homogenate. In addition, a kinetic method is indicated for the inhibition of an enzymatic reaction by a contaminant which binds the substrate to reduce the fraction available to the enzyme.  相似文献   

12.
1. The activities of the soluble reconstitutively active succinate dehydrogenase (EC 1.3.99.1) measured with three artificial electron acceptors, e.g. ferricyanide, phenazine methosulfate and free radical of N,N,N',N'-tetramethyl-p-phenylenediamine (WB), have been compared. The values estimated by extrapolation to infinite acceptor concentration using double reciprocal plots 1/v versus 1/[acceptor] are nearly the same for ferricyanide and phenazine methosulfate and about twice as high for the WB. 2. The double reciprocal plots 1/v versus 1/[succinate] in the presence of malonate at various concentrations of WB give a series of straight lines intercepting in the third quadrant. The data support the mechanism of the overall reaction, in which the reduced enzyme is oxidized by WB before dissociation of the enzyme-product complex. 3. The dependence of the rate of the overall reaction on WB concentration shows that only one kinetically significant redox site of the soluble succinate dehydrogenase is involved in the reduction of WB. 4. Studies of the change of V and Km values during aerobic inactivation of the soluble enzyme suggest that only 'the low Km ferricyanide reactive site' (Vinogradov, A.D., Gavrikova, E.V. and Goloveshkina, V.G. (1975) Biochem. Biophys, Res. Commun. 65, 1264--1269) is involved in reoxidation of the reduced enzyme by WB. 5. The pH dependence of V for the succinate-WB reductase reaction shows that the group of the enzyme with the pKa value of 6.7 at 22 degrees C is responsible for the reduction of dehydrogenase in the enzyme-substrate complex. 6. When WB interacts with the succinate-ubiquinone region of the respiratory chain, the double reciprocal plot 1/v versus 1/[WB] gives a straight line. The thenoyltrifluoroacetone inhibition of succinate-ubiquinone reductase or extraction of ubiquinone alter the 1/v versus 1/[WB] plots for the curves with a positive initial slope intercepting the ordinate at the same V as in the native particles. The data support the mechanism of succinate-ubiquinone reduction, in which no positive modulation of succinate dehydrogenase by ubiquinone exist in the membrane.  相似文献   

13.
The experimentally observed steady-level distribution of Na+ (25 degrees) and of D-glucose (0 degree c) in frog muscle were chosen as examples of solute distribution patterns observed in living cells, for comparison with those predicted by two theoretical models: one derived from the membrane-pump theory and the other from the association-induction (AI) hypothesis. Neither the distribution of Na+ nor that of D-glucose follows the pattern predicted by the membrane-pump models for solutes maintained at lower level than in the external medium, in which the plot of intracellular solute concentration as ordinate against different external concentrations as abscissa bends upward with increasing external solute concentration. Instead, both Na+ and D-glucose exhibit either straight line distribution with unchanging (below unity) slopes, or that of a hyperbola superimposed on such a straight line, both in agreement with the AI hypothesis.  相似文献   

14.
In a system derived from Escherichia coli, we carried out a detailed kinetic analysis of the inhibition of the puromycin reaction by lincomycin. N-Acetylphenylalanyl-tRNA (Ac-Phe-tRNA; the donor) reacts with excess puromycin (S) according to reaction [1], C+S Ks <--> CS k3 --> C'+P, where C is the Ac-Phe-tRNA-poly(U)-ribosome ternary complex (complex C). The entire course of reaction [1] appears as a straight line when the reaction is analyzed as pseudo-first-order and the data are plotted in a logarithmic form (logarithmic time plot). The slope of this straight line gives the apparent ksobs = k3[S]/(Ks + [S]). In the presence of lincomycin the logarithmic time plot is not a straight line, but becomes biphasic, giving an early slope (ke = k3[S]/(Ks(1 + [I]/Ki) + [S])) and a late slope (k1 = k3[S]/(Ks(1 + [I]/K'i + [S])). Kinetic analysis of the early slopes at various concentrations of S and I shows competitive inhibition with Ki = 10.0 microM. The late slopes also give competitive inhibition with a distinct inhibition constant K'i = 2.0 microM. Excluding alternative models, the two phases of inhibition are compatible with a model in which reaction [1] is coupled with reaction [2], C+I k4 <--> k5 CI k6 <--> k7 C*I, where the isomerization step CI <--> CI* is slower than the first step C+I <--> CI, Ki = k5/k4 and K'i = Ki [k7/(k6 + k7)]. Corroborative evidence for this model comes from the examination of reaction [2] alone in the absence of S. This reaction is analyzed as pseudo-first-order going toward equilibrium with kIeq = k7 + (k6 [I]/(Ki + [I])). The plot of kIeq versus [I] is not linear. This plot supports the two-step mechanism of reaction [2] in which k6 = 5.2 min-1 and k7 = 1.3 min-1. This is the first example of slow-onset inhibition of ribosomal peptidyltransferase which follows a simple model leading to the determination of the isomerization constants k6 and k7. We suggest that lincomycin inhibits protein synthesis by binding initially to the ribosome in competition with aminoacyl-tRNA. Subsequently, as a result of a conformational change, an isomerization occurs (CI <--> C*I), after which lincomycin continues to interfere with the binding of aminoacyl-tRNA to the isomerized complex.  相似文献   

15.
The effect of temperature and pH on kinetic behavior of α-galactosidase of Mortierella vinacea was investigated on the hydrolysis of p-nitrophenyl-α-D -galactopyranoside (PNPG). A very unusual kinetic behavior was observed for the soluble α-galactosidase i.e., substrate inhibition diminished gradually with increasing temperature or near the neutral pH range, and the kinetics approached the ordinary Michaelis-Menten (MM) type. On the other hand, with decreasing temperature or in acidic pH range, substrate inhibition was accelerated. Therefore, Arrhenius plots based on the initial reaction rate did not give straight lines. Furthermore, the slope in the Arrhenius plot changed with substrate concentration, which would make the determination of a characteristic value using conventional methods meaningless. However, the Arrhenius plots of individual kinetic parameters in the rate equation resulted in straight lines in the temperature range 15 to 50°C. From this, the drastic change in kinetic behavior could be explained in connection with the temperature and pH dependence of kinetic parameters in the model. For mold pellets (whole-cell enzyme), however, the influence of temperature and pH was less apparent than that of soluble enzyme because of the limitation in intraparticle diffusion. By using the rate equation that was determined for soluble enzyme and the theoretically derived effectiveness factor, the overall reaction rate for mold pellets at various temperature and pH could be predicted to some extent.  相似文献   

16.
Steady-state kinetic studies including initial velocity for mannitol oxidation and fructose reduction and product inhibition for mannitol oxidation using fructose and reduced nicotinamide adenine dinucleotide (NADH) are in accord with a reaction mechanism best described as ordered Bi-Bi with NAD+ and NADH designated as the first substrate, last product, respectively at pH 8.8. All replots of slopes and intercepts from product inhibition studies were linear. Dead-end inhibition studies using mannitol 1-phosphate gave slope-parabolic, intercept-linear noncompetitive inhibition for both NAD+ and mannitol as substrates. The dead-end inhibitor is capable of binding multiply to the E, EA, and EQ forms of the enzyme to an extent that is controlled by the concentration of substrates. The EQ complex is inferred to undergo a conformational change, E'Q equilibrium EQ, since (V1/E1) greater than (KiqV2)/(KqE1), and no evidence for dead-end complex formation with NADH can be adduced. This is interpreted to mean that the release of fructose from the central complex is faster than the isomerization of the E-NADH complex. When mannitol is saturating, the noncompetitive inhibition against NAD+, as the variable substrate, becomes parabolic uncompetitive. A replot of the slopes of the parabola against mannitol 1-phosphate remains concave upward. This situation could arise if the conformational change we infer in the EQ complex opens up additional sites on the protein which can interact with the dead-end inhibitor.  相似文献   

17.
Aspartate transcarbamylase (EC 2.1.3.2) catalyzes the bi substrate reaction—carbamyl phosphate+ L-aspartate ? carbamyl aspartate ? phosphate, The order of addition of substrates and release of products for the homogeneous aspartate transcarbamylase fromPhaseolus aureuss eedlings has been investigated by using the kinetic methods of analysis. p ]Initial velocity studies indicated that the mechanism might be a sequential one. Product inhibition studies showed that phosphate was a linear competitive inhibitor with respect to carbamyl phosphate and was anS (slope) andI (intercept) linear noncompetitive inhibitor with respect to aspartate. Carbamyl aspartate was a noncompetitive inhibitor with respect to both the substrates. These inhibition patterns agreed with an ordered mechanism of reaction with carbamyl phosphate as the leading substrate and phosphate as the last product to leave the enzyme surface. The presence of dead end complexes and the rapid equilibrium random mechanism were ruled out by the absence of inhibition by the substrate(s) and the linear replot slopevs. the inhibitor concentration. Acetyl phosphate, an analog ue of carbamyl phosphate was a non-competitive inhibitor with respect to aspartate. This result could be explained both in terms of an ordered as well as a random mechanism. On the other hand, succinate, an analog ue of aspartate was an uncompetitive inhibitor with respect to carbamyl phosphate, indicating that the mechanism was ordered. p ]The transition state analog ue, N-(phosphonoacetyl)-L-aspartate, binds much more tightly than either of the two substrates. This analog ue was a linear competitive inhibitor with respect to carbamyl phosphate and a linear noncompetitive inhibitor with respect to aspartate. These results are compatible with an ordered mechanism rather than a random one.  相似文献   

18.
The interaction of human organ alkaline phosphatases (orthophosphoric-monoester phosphohydrolases (alkaline optimum), EC 3.1.3.1) with sugars was studied. Hexosamines, N-acetylneuraminic acid (NANA or sialic acid), N-acetylmuramic acid and N-acetylglycolylneuraminic acid inhibited human organ alkaline phosphatase activities. Of these, sialic acid was the most effective inhibitor. The pH profiles for the enzymes in the absence and presence of sialic acid were similar. The sialic acid - enzyme complex was more heat stable than the free enzyme between 20 and 45 degrees C. Lineweaver-Burk plots of 1/v versus 1/S at various concentrations of sialic acid showed intersecting straight lines indicating that the mechanism of inhibition was a mixed type. The Ki value obtained from the plots of 1/v versus the square of sialic acid concentration was 0.07 mM for the hepatic, sialidase-treated hepatic, and intestinal alkaline phosphatases. The respective Hill coefficients varied somewhat with the alkaline phosphatase isoenzyme. Hyperbolic curves were obtained when the percentage of remaining activity was plotted against the substrate concentration at different concentrations of sialic acid. The Hill coefficient was lowered in the presence of sialic acid. The sialidase-treated hepatic enzymes used gave the most effective conversion. Partial denaturation of the enzyme with urea, or pronase digestion had a little if any effect on the sialic acid inhibition with constant time.  相似文献   

19.
S L Harbeson  D H Rich 《Biochemistry》1988,27(19):7301-7310
The synthesis and inhibition kinetics of a new, potent inhibitor of arginine aminopeptidase (aminopeptidase B; EC 3.4.11.6) are reported. The inhibitor is a reduced isostere of bestatin in which the amide carbonyl is replaced by the methylene (-CH2-) moiety. Analysis of the inhibition of arginine aminopeptidase by this inhibitor according to the method of Lineweaver and Burk yields an unusual noncompetitive double-reciprocal plot. The replot of the slopes versus [inhibitor] is linear (Kis = 66 nM), but the replot of the y intercepts (1/V) versus [inhibitor] is hyperbolic (Kii = 10 nM, Kid = 17 nM). These results provide evidence for a kinetic mechanism in which the inhibitor binds to the S1' and S2' subsites on the enzyme, not the S1 and S1' subsites occupied by dipeptide substrates. Furthermore, structure-activity data for a series of ketomethylene dipeptide isosteres in which the amide (-CONH-) of a dipeptide is replaced with the ketomethylene (-COCH2-) moiety show that the S1 and S1' subsites preferentially bind basic and aromatic side chains, respectively. These results are in agreement with the known substrate specificity of arginine aminopeptidase. The structure-activity data for several bestatin analogues, however, show that these compounds do not bind to the S1 and S1' sites of arginine aminopeptidase. A comparison of the data provides evidence that bestatin inhibits arginine aminopeptidase and possibly other aminopeptidases by binding to the S1' and S2' sites of the enzyme.  相似文献   

20.
Phosphatidic acid synthesis in Escherichia coli   总被引:10,自引:4,他引:6       下载免费PDF全文
The kinetic properties of acyl-coenzyme A (CoA): l-alpha-glycerol-phosphate trans-acylase (EC 2.3.1.15) from Escherichia coli were studied. At 10 C, a temperature at which the reaction was proportional to time and enzyme concentration, the enzyme had an apparent K(m) of 60 mum for l-alpha-glycerol-phosphate. The curve describing the velocity of the reaction as a function of palmitoyl-CoA concentration was sigmoid but the plot of v(-1) versus [S](-3) gave a straight line. A K(m) of about 11 mum was calculated for palmitoyl-CoA. Adenosine triphosphate specifically inhibited the reaction, being a noncompetitive inhibitor in respect to l-alpha-glycerol phosphate. Inhibition only occurred with high concentrations of palmitoyl-CoA, and maximal inhibition was 60%.  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