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 15 毫秒
1.
Reduced 3-thionicotinamide--adenine dinucleotide (sNADH) is shown to be fluorescent, with an emission maximum at 510 nm when excited in the region of the absorption maximum (398 nm), and with a very low quantum yield, (3.4 +/- 0.5) x 10(-4). The interaction between sNADH and octopine dehydrogenase was investigated by ultraviolet-difference spectroscopy and fluorescence. Some surprising fluorescence features were found when sNADH was bound to the enzyme in the presence of D-octopine, as follows. (a) There is an unusually high enhancement of the dinucleotide fluorescence (by at least a factor of 100) attended by a 40-nm blue shift of the emission maximum. (b) The protein fluorescence is quenched almost completely. (c) The bound coenzyme analog undergoes a photoreaction, which proceeds differently from that occurring the free form. These features appear to be unique to the octopine.sNADH complex, as for example they are not present when sNADH is bound to horse liver alcohol dehydrogenase, or when NADH is bound to octopine dehydrogenase. The possible origin of these fluorescence features is discussed. Binding and kinetic studies were carried out with sNAD and sNADH. It was found that sNAD neither binds nor acts kinetically as a coenzyme. sNADH exhibits relatively good binding, with Km and Ki values close to those of the natural coenzyme, but the turnover number is 460 times smaller than that with NADH. The kinetic consequences of these findings are discussed. The sNADH dissociation constants were determined as a function of temperature, and appear to be practically temperature-independent in the range 10--40 degrees C. It seems thus, in agreement with previous studies, that the interaction between octopine dehydrogenase and coenzymes proceeds athermically, regardless of the structure, affinity, and chemical reactivity of the coenzyme. The possible biological and chemical meaning of this finding is discussed.  相似文献   

2.
We investigated the temperature dependence of several functions of octopine dehydrogenase, a monomeric enzyme extracted from the shell fish Pecten maximus L. We found that six enzymatic functions are temperature independent or change only negligibly with temperatue. These are the dissociation constants of three coenzyme complexes and the Michaelis Km values for NAD, NADH and one of the substrates (D-octopine). This is taken as an indication of a temperature-regulatory mechanism which enables the enzyme to maintain a constant level of NAD, NADH and D-octopine in binary and ternary complexes independent of fluctuations of the external temperature. This is discussed with reference to enzymes from other poikilotherms, which reportedly display similar biologically meaningful response to temperature. We also discuss the meaning of our data from a thermodynamic viewpoint. Considering that in a temperature-independent binding process only entropy changes contribute to the standard free-energy change, we speculate on possible molecular models which might account for our results. We also investigate the activation-energy parameters for the reaction catalyzed by octopine dehydrogenase, as obtained from the temperature dependence of V. It is found that octopine dehydrogenase, relative to other dehydrogenases, is provided with a rather low delta H not equal to, which enables the enzyme to change its turnover number by only a small factor in the temperature range 5--35 degrees C.  相似文献   

3.
The kinetic scheme of octopine dehydrogenase of Pecten maximus L., a monomeric enzyme obeying a bi-ter sequential mechanism, was completed, essentially in the forward reaction, by steady-state studies over a wide range of substrate concentration at pH 7.0. Deviation from the Michaelis-Menten behavior with respect to NAD+ and other significant kinetic data led us to ascribe for octopine dehydrogenase mechanism the mnemonical enzyme concept. In addition, another regulatory behavior can be envisaged involving the formation of two dead-end complexes enzyme.NADH.D-octopine and enzyme.NAD+.pyruvate.L-arginine.  相似文献   

4.
5.
Aldehyde dehydrogenase 2 (ALDH2) catalyzes oxidation of toxic aldehydes to carboxylic acids. Physiologic levels of Mg(2+) ions influence ALDH2 activity in part by increasing NADH binding affinity. Traditional fluorescence measurements monitor the blue shift of the NADH fluorescence spectrum to study ALDH2-NADH interactions. By using time-resolved fluorescence spectroscopy, we have resolved the fluorescent lifetimes (τ) of free NADH (τ=0.4 ns) and bound NADH (τ=6.0 ns). We used this technique to investigate the effects of Mg(2+) on the ALDH2-NADH binding characteristics and enzyme catalysis. From the resolved free and bound NADH fluorescence signatures, the K(D) for NADH with ALDH2 ranged from 468 μM to 12 μM for Mg(2+) ion concentrations of 20 to 6000 μM, respectively. The rate constant for dissociation of the enzyme-NADH complex ranged from 0.4s(-1) (6000 μM Mg(2+)) to 8.3s(-1) (0 μM Mg(2+)) as determined by addition of excess NAD(+) to prevent re-association of NADH and resolving the real-time NADH fluorescence signal. The apparent NADH association/re-association rate constants were approximately 0.04 μM(-1)s(-1) over the entire Mg(2+) ion concentration range and demonstrate that Mg(2+) ions slow the release of NADH from the enzyme rather than promoting its re-association. We applied NADH fluorescence lifetime analysis to the study of NADH binding during enzyme catalysis. Our fluorescence lifetime analysis confirmed complex behavior of the enzyme activity as a function of Mg(2+) concentration. Importantly, we observed no pre-steady state burst of NADH formation. Furthermore, we observed distinct fluorescence signatures from multiple ALDH2-NADH complexes corresponding to free NADH, enzyme-bound NADH, and, potentially, an abortive NADH-enzyme-propanal complex (τ=11.2 ns).  相似文献   

6.
This work reports on the interaction of the fluorescent nicotinamide 1,N6-ethenoadenine dinucleotide (epsilonNAD+) with horse liver alcohol dehydrogenase, octopine dehydrogenase, and glyceraldehyde-3-phosphate dehydrogenase from different sources (yeast, lobster muscle, and rabbit muscle). The coenzyme fluorescence is enhanced by a factor of 10-13 in all systems investigated. It is shown that this enhancement cannot be due to changes in the polarity of the environment upon binding, and that it must be rather ascribed to structural properties of the bound coenzyme. Although dynamic factors could also be important for inducing changes in the quantum yield of epsilonNAD+ fluorescence, the close similarity of the fluorescence enhancement factor in all cases investigated indicates that the conformation of bound coenzyme is rather invariant in the different enzyme systems and overwhelmingly shifted toward an open form. Dissociation constants for epsilonNAD+-dehydrogenases complexes can be determined by monitoring the coenzyme fluorescence enhancement or the protein fluorescence quenching. In the case of yeast glyceraldehyde-3-phosphate dehydrogenase at pH 7.0 and t = 20 degrees the binding plots obtained by the two methods are coincident, and show no cooperativity. The affinity of epsilonNAD+ is generally lower than that of NAD+, although epsilonNAD+ maintains most of the binding characteristics of NAD+. For example, it forms a tight complex with horse liver alcohol dehydrogenase and pyrazole, and with octopine dehydrogenase saturated by L-arginine and pyruvate. One major difference in the binding behavior of NAD+ and epsilonNAD+ seems to be present in the muscle glyceraldehyde-3-phosphate dehydrogenase. In fact, no difference was found for epsilon NAD+ between the affinities of the third and fourth binding sites. The results and implications of this work are compared with those obtained recently by other authors.  相似文献   

7.
L Zeiri  E Reisler 《Biopolymers》1979,18(9):2289-2301
The CD of GDH–NADH complexes was measured in order to reexamine the binding of coenzyme to GDH. The existence of two distinct Cotton effects associated with two separate NADH binding sites/subunit was confirmed with native, polymerizing and crosslinked, unpolymerizing enzyme. CD titration of the high-affinity NADH sites revealed significant dependence of the optical activity of the bound coenzyme on the state of protein association. Molar ellipticity of bound NADH decreased with the increasing degree of polymerization of GDH. It is suggested that the high-affinity NADH sites are loacted at or near the association interface. Binding of NADH to the low-affinity sites, in the presence of GTP, leads to an inversion of the CD spectrum of GDH–NADH complexes. This inversion is not related to the polymerization of GDH. However, for proper analysis of the CD of NADH bound to the low-affinity sites, a correction for the effect of polymerization on the optical activity of NADH bound to the high-affinity sites is required.  相似文献   

8.
The fluorescence lifetime of nicotinamide adenine dinucleotide (NADH) is commonly used in conjunction with the phasor approach as a molecular biomarker to provide information on cellular metabolism of autofluorescence imaging of cells and tissue. However, in the phasor approach, the bound and free lifetime defining the phasor metabolic trajectory is a subject of debate. The fluorescence lifetime of NADH increases when bound to an enzyme, in contrast to the short multiexponential lifetime displayed by NADH in solution. The extent of fluorescence lifetime increase depends on the enzyme to which NADH is bound. With proper preparation of lactate dehydrogenase (LDH) using oxalic acid (OA) as an allosteric factor, bound NADH to LDH has a lifetime of 3.4 ns and is positioned on the universal semicircle of the phasor plot, inferring a monoexponential lifetime for this species. Surprisingly, measurements in the cellular environments with different metabolic states show a linear trajectory between free NADH at about 0.37 ns and bound NADH at 3.4 ns. These observations support that in a cellular environment, a 3.4 ns value could be used for bound NADH lifetime. The phasor analysis of many cell types shows a linear combination of fractional contributions of free and bound species NADH.  相似文献   

9.
Global analysis of fluorescence and associated anisotropy decays of intrinsic tissue fluorescence offers a sensitive and non-invasive probe of the metabolically critical free/enzyme-bound states of intracellular NADH in neural tissue. Using this technique, we demonstrate that the response of NADH to the metabolic transition from normoxia to hypoxia is more complex than a simple increase in NADH concentration. The concentration of free NADH, and that of an enzyme bound form with a relatively low lifetime, increases preferentially over that of other enzyme bound NADH species. Concomitantly, the intracellular viscosity is reduced, likely due to the osmotic swelling of mitochondria. These conformation and environmental changes effectively decrease the tissue fluorescence average lifetime, causing the usual total fluorescence increase measurements to significantly underestimate the calculated concentration increase. This new discrimination of changes in NADH concentration, conformation, and environment provides the foundation for quantitative functional imaging of neural energy metabolism.  相似文献   

10.
Deoxyhypusine is a modified lysine residue. It is formed posttranslationally in the precursor of eukaryotic initiation factor 5A (eIF5A) by deoxyhypusine synthase, employing spermidine as a butylamine donor. In the initial step of this reaction, deoxyhypusine synthase catalyzes the production of NADH through dehydrogenation of spermidine. Fluorescence measurements of this reaction revealed a -22-nm blue shift in the emission peak of NADH and a approximately 15-fold increase in peak intensity, characteristics of tightly bound NADH that were not seen by simply mixing NADH and enzyme. The fluorescent properties of the bound NADH can be ascribed to a hydrophobic environment and a rigidly held, open conformation of NADH, features in accord with the known crystal structure of the enzyme. Considerable fluorescence resonance energy transfer from tryptophan 327 in the active site to the dihydronicotinamide ring of NADH was seen. Upon addition of the eIF5A precursor, utilization of the enzyme-bound NADH for reduction of the eIF5A-imine intermediate to deoxyhypusine was reflected by a rapid decrease in the NADH fluorescence, indicating a transient hydride transfer mechanism as an integral part of the reaction. The number of NADH molecules bound approached four/enzyme tetramer; not all of the bound NADH was available for reduction of the eIF5A-imine intermediate.  相似文献   

11.
Binding of nicotinamide–adenine dinucleotides to diphtheria toxin   总被引:5,自引:0,他引:5       下载免费PDF全文
1. Changes in protein fluorescence have been utilized in determining the stoicheiometry and dissociation constants of the complexes of diphtheria toxin with NADH(2), NAD, NADPH(2) and NADP. 2. The binding stoicheiometry is 2moles of NADH(2) and 1mole of NADPH(2)/mole of diphtheria toxin. The binding sites for NADH(2) appear to be equivalent and independent. 3. The toxin shows a higher affinity for the reduced than for the oxidized forms of the nucleotides. 4. Dissociation constants at 0.01I, pH7 and 25 degrees are 0.7x10(-6)m for NADH(2) and 0.45x10(-6)m for NADPH(2). Dissociation constants increase with increasing ionic strength, indicating that the binding is mainly electrostatic. 5. Bound NADH(2) and NADPH(2) may be activated to fluoresce by the transfer of energy from the excited aromatic amino acids of the toxin. Activation and emission spectra of bound and free nucleotides are compared. 6. Since NAD and NADH(2) are cofactors specifically required for the inhibition of protein synthesis by diphtheria toxin, the possible role of toxin-nucleotide complexes is discussed in this regard.  相似文献   

12.
The techniques of fluorescence enhancement, fluorescence quenching, fluorescence polarization, and equilibrium dialysis are utilized to study the binding properties of coenzyme to horse liver alcohol dehydrogenase. Polarization of fluorescence and equilibrium dialysis show that NADH binds to alcohol dehydrogenase with a stoichiometry of 6 mol per mol of enzyme, in contrast to the value of 2 determined from fluorescence enhancement measurements. NAD+ also binds with a stoichiometry of six as was determined by equilibrium dialysis. The two NADH sites which bind coenzyme more tightly and which are revealed by fluorescence enhancement measurements are designated the catalytic sites. Binding of coenzyme to the four ancillary sites does not alter the quantum yield of NADH but results in a 20% contribution to quenching of enzyme's tryptophan fluorescence. From the emission anisotropy of bound NADH of 24.0% for the additional sites and 28.1% for the catalytic sites and their relative fluorescence lifetimes at the same wavelengths of excitation and emmision, we conclude that the nicotinamide ring of NADH bound to the additional sites exhibits a freedom of motion independent of the macromolecule, while that bound to the catalytic sites is more rigidly held. Polarization of fluorescence yields negative intrinsic free energies of 9.2 and 7.5 Cal M-1 for NADH interaction with the catalytic and additional sites, respectively. Although these values are 1.3 to 2.0 Cal higher than those determined by fluorescence quenching and equilibrium dialysis, the mean Hill coefficient of 1.76 plus or minus 0.06, the titration span of 2.4 logarithmic units and coupling free energies (in magnitude and sign) are the same for all these techniques. The above difference in the intrinsic free energies are attributed largely to the different modes of interaction of excited and unexcited NADH molecules with alcohol dehydrogenase.  相似文献   

13.
Two octopine dehydrogenases in crown-gall tumor tissue   总被引:1,自引:0,他引:1  
Extracts from four crown-gall tumor tissue culture lines, originally induced by two octopine-type strains of Agrobacterium on three plant species, converted l-arginine-[5-3H] to a compound which co-migrated with octopine on electrophoresis. Synthesis showed dependence on added pyruvate and reduced pyridine nucleotide. Both NADH and NADPH were active and mixtures of the two coenzymes, when tested with Vinca strain W1 tumor extracts, were more effective than either coenzyme at comparable concentrations. Addition of an NADH-consuming enzyme system to reaction mixtures containing NADPH had little effect on this activity. Products formed by Vinca rosea strain W1 tumor extracts and Phaseolus vulgaris strain B6 tumor extracts in reaction mixtures containing pyruvate plus NADH or NADPH co-eluted with unlabeled octopine on ion exchange chromatography. The product from the Vinca reaction mixtures co-migrated with an octopine standard in three TLC systems. Permanganate treatment of the enzymatically formed tritiated product and of unlabeled octopine gave compounds with Rf, similar to arginine and γ-guanidinobutyric acid, the products expected from permanganate degradation of octopine. The Vinca W1 extracts catalyzed the oxidative cleavage of octopine, with the formation of arginine, in the presence of NAD or NADP. Two octopine dehydrogenases were concluded to be present in these tissues, one dependent on NAD, the second on NADP.  相似文献   

14.
R Koren  G G Hammes 《Biochemistry》1975,14(5):1021-1025
The interaction of NADH with s-malate dehydrogenase isolated from beef heart was studied in 20 mM potassium phosphate (pH 6.9)-1 mM EDTA, with forced dialysis, fluorescence, and temperature-jump techniques. Measurements of the change in fluorescence of NADH when it is titrated with enzyme indicate NADH bound to monomeric and dimeric enzyme have different fluorescence yields. These data and the results of direct binding studies can be explained in terms of a model in which the NADH binding sites on dimeric enzyme are equivalent or nearly equivalent, and NADH binding to monomeric enzyme occurs with an affinity very similar to that of the dimer. However, the fluorescence enhancement of NADH on binding to the enzyme is different for the monomer and for each of the two dimer sites.  相似文献   

15.
The formation of binary complexes between sturgeon apoglyceralddhyde-3-phosphate dehydrogenase, coenzymes (NAD+ and NADH) and substrates (phosphate, glyceraldehyde 3-phosphate and 1,3-bisphosphoglycerate) has been studied spectrophotometrically and spectrofluorometrica-ly. Coenzyme binding to the apoenzyme can be characterized by several distinct spectroscopic properties: (a) the low intensity absorption band centered at 360 nm which is specific of NAD+ binding (Racker band); (b) the quenching of the enzyme fluorescence upon coenzyme binding; (c) the quenching of the fluorescence of the dihydronicotinamide moiety of the reduced coenzyme (NADH); (D) the hypochromicity and the red shift of the absorption band of NADH centered at 338 nm; (e) the coenzyme-induced difference spectra in the enzyme absorbance region. The analysis of these spectroscopic properties shows that up to four molecules of coenzyme are bound per molecule of enzyme tetramer. In every case, each successively bound coenzyme molecule contributes identically to the total observed change. Two classes of binding sites are apparent at lower temperatures for NAD+ Binding. Similarly, the binding of NADH seems to involve two distinct classes of binding sites. The excitation fluorescence spectra of NADH in the binary complex shows a component centered at 260 nm as in aqueous solution. This is consistent with a "folded" conformation of the reduced coenzyme in the binary complex, contradictory to crystallographic results. Possible reasons for this discrepancy are discussed. Binding of phosphorylated substrates and orthophosphate induce similar difference spectra in the enzyme absorbance region. No anticooperativity is detectable in the binding of glyceraldehyde 3-phosphate. These results are discussed in light of recent crystallographic studies on glyceraldehyde-3-phosphate dehydrogenases.  相似文献   

16.
NADH is a naturally fluorescent metabolite associated with cellular respiration. Exploiting the different fluorescence lifetime of free and bound NADH has the potential to quantify the relative amount of bound and free NADH, enhancing understanding of cellular processes including apoptosis, cancer pathology, and enzyme kinetics. We use the phasor-fluorescence lifetime image microscopy approach to spatially map NADH in both the free and bound forms of live undifferentiated and differentiated myoblast cells. The phasor approach graphically depicts the change in lifetime at a pixel level without the requirement for fitting the decay. Comparison of the spatial distribution of NADH in the nucleus of cells induced to differentiate through serum starvation and undifferentiated cells show differing distributions of bound and free NADH. Undifferentiated cells displayed a short lifetime indicative of free NADH in the nucleus and a longer lifetime attributed to the presence of bound NADH outside of the nucleus. Differentiating cells displayed redistribution of free NADH with decreased relative concentration of free NADH within the nucleus whereas the majority of NADH was found in the cytoplasm.  相似文献   

17.
The binding of oxidized and reduced coenzyme (NAD+ and NADH) to 3-phosphoglyceroyl-glyceraldehyde-3-phosphate dehydrogenase has been studied spectrophotometrically and fluorimetrically. The binding of NAD+ to the acylated sturgeon enzyme is characterized by a significant quenching of the enzyme fluorescence (about 25%) and the induction of a difference spectrum in the ultraviolet absorbance region of the enzyme. Both of these spectroscopic properties are quantitatively distinguishable from those of the corresponding binary enzyme-NAD+ complex. Binding isotherms estimated by gel filtration of the acylated enzyme are in close agreement to those obtained by spectrophotometric and fluorimetric titrations. Up to four NAD+ molecules are bound to the enzyme tetramer. No anticooperativity can be detected in the binding of oxidized coenzyme, which is well described on the basis of a single class of four binding sites with a dissociation constant of 25 muM at 10 degrees C, pH 7.0. The binding of NADH to the acylenzyme has been characterized spectrophotometrically. The absorption band of the dihydronicotinamide moiety of the coenzyme is blue-shifted to 335 nm with respect to free NADH. In addition, a large hypochromicity (23%) is observed together with a significant increase of the bandwidth at half height of this absorption band. This last property is specific to the acylenzyme-DADH complex, since it disappears upon arsenolysis of the acylenzyme. The binding affinity of NADH to the acylated enzyme has been estimated by performing simultaneous spectrophotometric and fluorimetric titrations of the NADH appearance upon addition of NAD+ to a mixture of enzyme and excess glyceraldehyde 3-phosphate. In contrast to NAD+, the reduced coenzyme NADH appears to be relatively strongly bound to the acylated enzyme, the dissociation constant of the acylenzyme-NADH complex being estimated as 2.0 muM at 25 degrees C. In addition a large quenching of the NADH fluorescence (about 83%) is observed. The comparison of the dissociation constants of the coenzyme-acylenzyme complexes and the corresponding Michaelis constants suggests a reaction mechanism of the enzyme in which significant formation and dissociation of NAD+-acylenzyme and NADH-acylenzyme complexes occur. Under physiological conditions the activity of the enzyme can be regulated by the ratio of oxidized and reduced coenzymes. Possible reasons for the lack of anticooperativity in coenzyme binding to the acylated form of the enzyme are discussed.  相似文献   

18.
The cellular proportion of free and protein‐bound NADH complexes is increasingly recognized as a metabolic indicator and biomarker. Because free and bound forms exhibit different fluorescence spectra, we consider whether autofluorescence shape sufficiently correlates with mitochondrial metabolism to be useful for monitoring in cellular suspensions. Several computational approaches for rapidly quantifying spectrum shape are used to detect Saccharomyces cereviseae response to oxygenation, and to the addition of mitochondrial functional modifiers and metabolic substrates. Observed changes appear consistent with previous studies probing free/protein‐bound proportions, making this a potentially useful approach for the real‐time monitoring of metabolism. (© 2015 WILEY‐VCH Verlag GmbH &Co. KGaA, Weinheim)  相似文献   

19.
Pd(II) complexes of two anthracyclines, adriamycin and daunorubicin, have been studied. Using potentiometric absorption, fluorescence, and circular dichroism measurements, we have shown that adriamycin can form two complexes with Pd(II). The first complex (I) involves two molecules of drug per Pd(II) ion; one of the molecules is chelated to Pd(II) through the carbonyl oxygen on C12 and the phenolate oxygen on C11, and the other one is bound to Pd(II) through the nitrogen of the amino sugar. This complexation induces a stacking of the two molecules of drug. In the second complex (II), two Pd(II) ions are bound to two molecules of drug (A1 and A2). One Pd(II) is bound to the oxygen on the carbons C11 and C12 of molecule A1 and the amino sugar of molecule A2 whereas the second Pd(II) ion is bound to the oxygen on C11 and C12 of molecule A2 and the amino sugar of molecule A1. The same complexes are formed between Pd(II) and daunorubicin. The stability constant for complex II is beta = (1.3 +/- 0.5) X 10(22). Interaction with DNA has been studied, showing that almost no modification of the complex occurred. This complex displays antitumor activity against P-388 leukemia that compares with that of the free drug. Complex II, unlike adriamycin, does not catalyze the flow of electrons from NADH to molecular oxygen through NADH dehydrogenase.  相似文献   

20.
The binding of the coenzyme to octopine dehydrogenase was investigated by kinetic and spectroscopic studies using different analogues of NAD+. The analogues employed were fragments of the coenzyme molecule and dinucleotides modified on the purine or the pyridine ring. The binding of ADPribose is sufficient to induce local conformational changes necessary for the good positioning of substrates. AMP, ADP, NMN+ and NMNH do not show this effect. Analogues modified on the purine ring such as nicotinamide deaminoadenine dinucleotide, nicotinamide--8-bromoadenine dinucleotide, nicotinamide--8-thioadenine dinucleotide and nicotinamide 1: N6-ethenoadenine dinucleotide bind to the enzyme and give catalytically active ternary complexes. Modifications of the pyridine ring show an important effect on the binding of the coenzyme as well as on the formation of ternary complexes. Thus, the carboxamide group can well be replaced by an acetyl group and also, though less efficiently, by a formyl or cyano group. However more bulky substituents such as thio, chloroacetyl or propionyl groups prevent the binding. The analogues bearing a methyl group in the 4 or 5 position, which are competitive inhibitors, are able to give binary by not ternary complexes. The case of 1,4,5,6-tetrahydronicotinamide--adenine dinucleotide which does not give ternary complexes like NADH is discussed. The above findings show that the pyridine and adenine parts are both involved in the binding of the coenzyme and of the substrate to octopine dehydrogenase. The nicotinamide binding site of this enzyme seems to be the most specific and restricted one among the dehydrogenases so far described. The protective effects of coenzyme analogues towards essential -SH group were also studied.  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