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.
A Glu141Asn mutant Paracoccus sp. 12-A formate dehydrogenase catalyzes marked glyoxylate reduction. Additional replacement of the His332-Gln313 pair with His-Glu, which is a consensus acid/base catalyst in D-hydroxyacid dehydrogenases, further improved the catalytic activity of the enzyme as to glyoxylate reduction through enhancement of the hydrogen transfer step in the catalytic process, slightly shifting the optimal pH for the reaction. On the other hand, the replacement induced no marked activity toward other 2-ketoacid substrates, and diminished the enzyme activity as to formate oxidation. Consequently, the formate dehydrogenase was converted to a highly specific and active glyoxylate reductase through only the two amino acid replacements.  相似文献   

2.
Activity of the mammalian pyruvate dehydrogenase complex is regulated by phosphorylation-dephosphorylation of three specific serine residues (site 1, Ser-264; site 2, Ser-271; site 3, Ser-203) of the alpha subunit of the pyruvate dehydrogenase (E1) component. Phosphorylation is carried out by four pyruvate dehydrogenase kinase (PDK) isoenzymes. Specificity of the four mammalian PDKs toward the three phosphorylation sites of E1 was investigated using the recombinant E1 mutant proteins with only one functional phosphorylation site present. All four PDKs phosphorylated site 1 and site 2, however, with different rates in phosphate buffer (for site 1, PDK2 > PDK4 approximately PDK1 > PDK3; for site 2, PDK3 > PDK4 > PDK2 > PDK1). Site 3 was phosphorylated by PDK1 only. The maximum activation by dihydrolipoamide acetyltransferase was demonstrated by PDK3. In the free form, all PDKs phosphorylated site 1, and PDK4 had the highest activity toward site 2. The activity of the four PDKs was stimulated to a different extent by the reduction and acetylation state of the lipoyl moieties of dihydrolipoamide acetyltransferase with the maximum stimulation of PDK2. Substitution of the site 1 serine with glutamate, which mimics phosphorylation-dependent inactivation of E1, did not affect phosphorylation of site 2 by four PDKs and of site 3 by PDK1. Site specificity for phosphorylation of four PDKs with unique tissue distribution could contribute to the tissue-specific regulation of the pyruvate dehydrogenase complex in normal and pathophysiological states.  相似文献   

3.
Unfertilized eggs and early embryos up to the 2-day (16-cell) cleavage stage of development in the rabbit contain predominantly the most cathodal lactate dehydrogenase isoenzyme made up of A-type subunits. Following early cleavage there is a progressing increase in total LDH activity in the embryo as development proceeds through 4- and 6-day blastocyst stages. This is accompanied by an increase in the amount of B-type subunits and a concomitant shift in the lactate dehydrogenase isoenzyme electrophoretic pattern toward the anodal isoenzyme types.  相似文献   

4.
Fibroblasts from patients with long-chain acyl-CoA dehydrogenase deficiency were found to oxidize [1-14C]linoleate at an average rate of 60% of normal but [9,10(n)-3H]myristate at an average rate of only 37% of normal, a relationship reverse from that predicted by the chain-length specificities of the three known straight-chain mitochondrial acyl-CoA dehydrogenases. The residual long-chain beta-oxidative activity was found to be mitochondrial and associated with the accumulation of tetradecadienoate (C14:2w6) when the mutant fibroblasts were incubated with 100 mumol/L linoleate (C18:2w6) or eicosadienoate (C20:2w6). The results suggest the presence in human fibroblasts of a novel acyl-CoA dehydrogenase with activity toward 15 to 20 carbon-length fatty acids.  相似文献   

5.
L-histidinol dehydrogenase, a Zn2+-metalloenzyme   总被引:3,自引:0,他引:3  
The enzymatic activity of L-histidinol dehydrogenase from Salmonella typhimurium was stimulated by the inclusion of 0.5 mM MnCl2 in the assay medium. At pH 9.2 the stimulation was correlated with binding of 1 g-atom of 54Mn2+/mol dimer, KD = 37 microM. ZnCl2, which prevented the MnCl2 stimulation, also bound to the enzyme, 1.2 g-atom/mol dimer, KD = 51 microM, and prevented Mn2+ binding. Enzyme activity was lost when histidinol dehydrogenase was incubated in 8 M urea. Reactivation was observed when urea-denatured enzyme was diluted into buffer containing 2-mercaptoethanol but required either MnCl2 or ZnCl2. Histidinol dehydrogenase was inactivated by the transition metal chelator 1,10-phenanthroline or by high levels of 2-mercaptoethanol. The nonchelating 1,7-phenanthroline was not an inactivator, and inactivation by either 1,10-phenanthroline or 2-mercaptoethanol was prevented by MnCl2. Enzyme inactivated by 1,10-phenanthroline could be reactivated by addition of MnCl2 or ZnCl2 in the presence of 2-mercaptoethanol. Reactivation was correlated with the binding of 1.5 g-atom 54Mn2+/mol dimer. Atomic absorption analysis of the native enzyme indicated the presence of 1.65 g-atom Zn/mol dimer, and no Mn was detected. The results demonstrate, therefore, that histidinol dehydrogenase contains two metal binding sites per enzyme dimer, which normally bind Zn2+, but which may bind Mn2+ while retaining enzyme function. Histidinol dehydrogenase is thus the third NAD-linked oxidoreductase in which Zn2+ fulfills an essential structural and/or catalytic role.  相似文献   

6.
The constitutive xanthine dehydrogenase and the inducible 2-furoyl-coenzyme A (CoA) dehydrogenase could be labeled with [185W]tungstate. This labeling was used as a reporter to purify both labile proteins. The radioactivity cochromatographed predominantly with the residual enzymatic activity of both enzymes during the first purification steps. Both radioactive proteins were separated and purified to homogeneity. Antibodies raised against the larger protein also exhibited cross-reactivity toward the second smaller protein and removed xanthine dehydrogenase and 2-furoyl-CoA dehydrogenase activity up to 80 and 60% from the supernatant of cell extracts, respectively. With use of cell extract, Western immunoblots showed only two bands which correlated exactly with the activity stains for both enzymes after native polyacrylamide gel electrophoresis. Molybdate was absolutely required for incorporation of 185W, formation of cross-reacting material, and enzymatic activity. The latter parameters showed a perfect correlation. This evidence proves that the radioactive proteins were actually xanthine dehydrogenase and 2-furoyl-CoA dehydrogenase. The apparent molecular weight of the native xanthine dehydrogenase was about 300,000, and that of 2-furoyl-CoA dehydrogenase was 150,000. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of both enzymes revealed two protein bands corresponding to molecular weights of 55,000 and 25,000. The xanthine dehydrogenase contained at least 1.6 mol of molybdenum, 0.9 ml of cytochrome b, 5.8 mol of iron, and 2.4 mol of labile sulfur per mol of enzyme. The composition of the 2-furoyl-CoA dehydrogenase seemed to be similar, although the stoichiometry was not determined. The oxidation of furfuryl alcohol to furfural and further to 2-furoic acid by Pseudomonas putida Fu1 was catalyzed by two different dehydrogenases.  相似文献   

7.
An acetaldehyde dehydrogenase from germinating seeds   总被引:2,自引:2,他引:0       下载免费PDF全文
An acetaldehyde dehydrogenase from germinating peanut cotyledons has been purified and its properties have been studied. At the highest purification achieved the preparation is free of alcohol dehydrogenase activity.

The enzyme is specific toward diphosphopyridine nucleotide, and can oxidize a variety of aldehydes. The highest reaction rate is obtained with acetaldehyde, which is oxidized to acetate. All the attempts to demonstrate the formation of an energy-rich acetyl derivative during the course of the reaction failed. The enzyme is inhibited by aldol; it is sensitive toward sulfhydryl reagents, including arsenite. Reduced glutathione stabilizes the enzyme, while cysteine, mercaptoethanol, and coenzyme A are inhibitory.

Acetaldehyde dehydrogenase is activated by phosphate and inhibited by fatty acyl-CoA derivatives. It appears to be activated by the substrate, as was deduced from the shape of the plot of reaction velocity against acetaldehyde. These properties suggest that the enzyme is an allosteric protein.

The plot of reaction velocity against substrate concentration is anomalous. The shape of this plot seems to reflect the presence of 2 different enzymatic activities, one with extremely high apparent affinity for acetaldehyde. The 2 activities may reflect 2 conformational states of a single enzyme or 2 separate enzymes.

Experiments with tissue slices indicate that the reaction catalyzed by this enzyme is a step in the oxidation of ethanol to acetyl-CoA. This enzyme may also participate in the oxidation of pyruvate to acetyl-CoA in certain tissues.

  相似文献   

8.
Two forms of NAD-dependent D-mandelate dehydrogenase (D-ManDHs) were purified from Enterococcus faecalis IAM 10071. While these two enzymes consistently exhibited high activity toward large 2-ketoacid substrates that were branched at the C3 or C4 position, they gave distinctly different K(m) and V(max) values for these substrates and had distinct molecular weights by gel electrophoresis and gel filtration.  相似文献   

9.
2-Oxoglutarate (2-OG)-dependent O2 uptake by washed or purified turnip (Brassica rapa L.) and pea (Pisum sativum L. cv. Massey Gem) leaf mitochondria, in the presence of malonate, was inhibited between 65 and 90% by micromolar levels of pyruvate. The inhibition was not observed in the absence of malonate and was reversed by alpha-cyano-4-hydroxycinnamic acid. The inhibition was also reversed by oxaloacetate or by malate, but not by any other tricarboxylic acid cycle intermediates. The stimulation of O2 uptake by oxaloacetate was half maximal at 8-9 microM and was transient, indicating its action was not mediated through the complete metabolic removal of pyruvate. Pyruvate had not effect on 2-OG oxidation under conditions in which pyruvate dehydrogenase was not active, indicating that pyruvate metabolism, rather than pyruvate itself, was responsible for producing the inhibition of 2-OG oxidation. Similar results were obtained with detergent-treated mitochondrial extracts with the exception that the inhibition of 2-OG oxidation by pyruvate could also be reversed by coenzyme A. The results suggest that pyruvate inhibits 2-oxoglutarate oxidation, in intact plant mitochondria, by sequestering intramitochondrial CoA as acetyl-CoA and, in the absence of citrate synthase activity, reduces the amount of free coenzyme A available for 2-oxoglutarate dehydrogenase. These results indicate that pyruvate dehydrogenase and 2-oxoglutarate dehydrogenase share a common CoA pool within plant mitochondria and that the turnover of the acyl-CoA product of one enzyme will dramatically influence the activity of the other.  相似文献   

10.
The regulatory effects of alpha-ketoisovalerate on purified bovine heart pyruvate dehydrogenase complex and endogenous pyruvate dehydrogenase kinase were investigated. Incubation of pyruvate dehydrogenase complex with 0.125 to 10 mM alpha-ketoisovalerate caused an initial lag in enzymatic activity, followed by a more linear but inhibited rate of NADH production. Incubation with 0.0125 or 0.05 mM alpha-ketoisovalerate caused pyruvate dehydrogenase inhibition, but did not cause the initial lag in pyruvate dehydrogenase activity. Gel electrophoresis and fluorography demonstrated the incorporation of acyl groups from alpha-keto[2-14C]isovalerate into the dihydrolipoyl transacetylase component of the enzyme complex. Acylation was prevented by pyruvate and by arsenite plus NADH. Endogenous pyruvate dehydrogenase kinase activity was stimulated specifically by K+, in contrast to previous reports, and kinase stimulation by K+ correlated with pyruvate dehydrogenase inactivation. Maximum kinase activity in the presence of K+ was inhibited 62% by 0.1 mM thiamin pyrophosphate, but was inhibited only 27% in the presence of 0.1 mM thiamin pyrophosphate and 0.1 mM alpha-ketoisovalerate. Pyruvate did not affect kinase inhibition by thiamin pyrophosphate at either 0.05 or 2 mM. The present study demonstrates that alpha-ketoisovalerate acylates heart pyruvate dehydrogenase complex and suggests that acylation prevents thiamin pyrophosphate-mediated kinase inhibition.  相似文献   

11.
Summary Partially purified flounder muscle (Pseudopleuronectus americanus) glyceraldehyde 3-phosphate dehydrogenase was immobilized on cyanogen bromide-activated Sepharose. The catalytic properties of the immobilized preparation were studied to determine if immobilization alters the kinetic properties of the native holoenzyme. The results indicate that the pH activity profile of immobilized glyceraldehyde 3-phosphate dehydrogenase did not differ from that of the native enzyme. The Michaelis constants (Km) for NAD and glyceraldehyde 3-phosphate were somewhat altered. The enzyme stability toward various inactivation treatments in the presence and absence of NAD was characterized and compared to that of he native enzyme. When either form of the enzyme was incubated with urea at concentrations greater than 2m, inactivation occurred very rapidly. Incubation in 0.1% trypsin for 60 minutes decreased the activity of immobilized glyceraldehyde 3-phosphate dehydrogenase by 45% and of the native soluble enzyme by 70%. The immobilized enzyme also exhibited considerably more stability than the native soluble enzyme when exposed to a temperature of 50° or to 20 mm ATP. In all cases NAD either greatly reduced the rate of inactivation or completely protected the enzyme from inactivation.  相似文献   

12.
Eight distinct forms of the soluble 17alpha-hydroxy steroid dehydrogenase of rabbit liver were resolved by DEAE-cellulose chromatography and isoelectric focusing. Five of these enzymes were homogeneous as judged by polyacrylamide-gel electrophoresis. Substrate-specificity studies carried out with oestradiol-17alpha and oestradiol-17alpha 3-glucuronide revealed a variation in activity toward these substrates among the different purified enzyme forms. Three forms of the 17alpha-hydroxy steroid dehydrogenase exhibited equal activity toward both oestrogen substrates, whereas three forms of the enzyme displayed a greater activity toward the glucuronide derivative of oestradiol-17alpha. One enzyme in particular is essentially specific for oestradiol-17alpha 3-glucuronide, its activity toward oestradiol-17alpha being only one-thirtieth that observed with the 3-glucuronide derivative.  相似文献   

13.
In a previous work, we presented evidence for the presence of a protein encoded by At5g50600 in oil bodies (OBs) from Arabidopsis thaliana [P. Jolivet, E. Roux, S. D'Andrea, M. Davanture, L. Negroni, M. Zivy, T. Chardot, Protein composition of oil bodies in Arabidopsis thaliana ecotype WS, Plant Physiol. Biochem. 42 (2004) 501-509]. Using specific antibodies and proteomic techniques, we presently confirm the existence of this protein, which is a member of the short-chain steroid dehydrogenase reductase superfamily. We have measured its activity toward various steroids (cholesterol, dehydroepiandrosterone, cortisol, corticosterone, estradiol, estrone) and NAD(P)(H), either within purified OBs or as a purified bacterially expressed chimera. Both enzymatic systems (OBs purified from A. thaliana seeds as well as the chimeric enzyme) exhibited hydroxysteroid dehydrogenase (HSD) activity toward estradiol (17beta-hydroxysteroid) with NAD+ or NADP+, NADP+ being the preferred cofactor. Low levels of activity were observed with cortisol or corticosterone (11beta-hydroxysteroids), but neither cholesterol nor DHEA (3beta-hydroxysteroids) were substrates, whatever the cofactor used. Similar activity profiles were found for both enzyme sources. Purified OBs were found to be also able to catalyze estrone reduction (17beta-ketosteroid reductase activity) with NADPH. The enzyme occurring in A. thaliana OBs can be classified as a NADP+-dependent 11beta-,17beta-hydroxysteroid dehydrogenase/17beta-ketosteroid reductase. This enzyme probably corresponds to AtHSD1, which is encoded by At5g50600. However, its physiological role and substrates still remain to be determined.  相似文献   

14.
Pseudomonads are the only organisms so far known to produce two lipoamide dehydrogenases (LPDs), LPD-Val and LPD-Glc. LPD-Val is the specific E3 component of branched-chain oxoacid dehydrogenase, and LPD-Glc is the E3 component of 2-ketoglutarate and possibly pyruvate dehydrogenases and the L-factor of the glycine oxidation system. Three mutants of Pseudomonas putida, JS348, JS350, and JS351, affected in lpdG, the gene encoding LPD-Glc, have been isolated; all lacked 2-ketoglutarate dehydrogenase, but two, JS348 and JS351, had normal pyruvate dehydrogenase activity. The pyruvate and 2-ketoglutarate dehydrogenases of the wild-type strain of P. putida were both inhibited by anti-LPD-Glc, but the pyruvate dehydrogenase of the lpdG mutants was not inhibited, suggesting that the mutant pyruvate dehydrogenase E3 component was different from that of the wild type. The lipoamide dehydrogenase present in one of the lpdG mutants, JS348, was isolated and characterized. This lipoamide dehydrogenase, provisionally named LPD-3, differed in molecular weight, amino acid composition, and N-terminal amino acid sequence from LPD-Glc and LPD-Val. LPD-3 was clearly a lipoamide dehydrogenase as opposed to a mercuric reductase or glutathione reductase. LPD-3 was about 60% as effective as LPD-Glc in restoring 2-ketoglutarate dehydrogenase activity and completely restored pyruvate dehydrogenase activity in JS350. These results suggest that LPD-3 is a lipoamide dehydrogenase associated with an unknown multienzyme complex which can replace LPD-Glc as the E3 component of pyruvate and 2-ketoglutarate dehydrogenases in lpdG mutants.  相似文献   

15.
A NADP(+)-dependent 3 beta-hydroxysteroid dehydrogenase activity was localized in the microsomal fraction of rat liver. This enzyme was solubilized and separated completely from 3 alpha-hydroxysteroid dehydrogenase by Matrex red A column chromatography. Partially purified 3 beta-hydroxysteroid dehydrogenase catalyzed the oxidation and reduction between the 3 beta-hydroxyl and 3-ketonic group of steroids or bile acids having no double bond in the A/B ring, but was inactive toward 3 alpha-hydroxyl group. The enzyme required NADP+ for oxidation and NADPH for reduction. The activity was inhibited by p-chloromercuribenzoic acid or p-chloromercuribenzenesulfonic acid at the concentration of 10(-4) M. The molecular weight of the enzyme was estimated to be about 43,000 by Sephadex G-200 column chromatography. From these results, it is concluded that the enzyme is a new type of microsomal NADP+:3 beta-hydroxysteroid dehydrogenase.  相似文献   

16.
The open reading frame TM1643 of Thermotoga maritima belongs to a large family of proteins, with homologues in bacteria, archaea, and eukaryotes. TM1643 is found in an operon with two other genes that encode enzymes involved in the biosynthesis of NAD. In several bacteria, the gene in the position occupied by TM1643 encodes an aspartate oxidase (NadB), which synthesizes iminoaspartate as a substrate for NadA, the next enzyme in the pathway. The amino acid sequence of TM1643 does not share any recognizable homology with aspartate oxidase or with other proteins of known functions or structures. To help define the biological functions of TM1643, we determined its crystal structure at 2.6A resolution and performed a series of screens for enzymatic function. The structure reveals the presence of an N-terminal Rossmann fold domain with a bound NAD(+) cofactor and a C-terminal alpha+beta domain. The structural information suggests that TM1643 may be a dehydrogenase and the active site of the enzyme is located at the interface between the two domains. The enzymatic characterization of TM1643 revealed that it possesses NAD or NADP-dependent dehydrogenase activity toward l-aspartate but no aspartate oxidase activity. The product of the aspartate dehydrogenase activity is also iminoaspartate. Therefore, our studies demonstrate that two different enzymes, an oxidase and a dehydrogenase, may have evolved to catalyze the first step of NAD biosynthesis in prokaryotes. TM1643 establishes a new class of amino acid dehydrogenases.  相似文献   

17.
Methanol dehydrogenase was found to be present in subcellular preparations of methanol-grown Methylosinus trichosporium and occurred almost wholly in the soluble fraction of the cell. The enzyme, purified by DEAE-Sephadex and Sephadex G-100 chromatography, showed broad specificity toward different substrates and oxidized the aromatic alcohols benzyl, vanillyl, and veratryl alcohols in addition to a range of aliphatic primary alcohols. No enzyme activity was found toward the corresponding aldehydes of the alcohols tested. The Km for methanol was 50 microM, and that for the aromatic alcohols was in the range of 1 to 2 mM. EDTA and p-nitrophenylhydrazine, which are inhibitors of methanol oxidation in whole cells of methylotrophs, had little effect on activity of the purified enzyme. The results now extend the range of substrates oxidized by methanol dehydrogenase to include the aromatic alcohols.  相似文献   

18.
Rat liver alcohol dehydrogenase was purified and four isoenzyme forms, demonstrated by starch gel electrophoresis, were separated by O-(carboxymethyl)-cellulose chromatography. Each of the isoenzymes had a distinct isoelectric point. All isoenzymes were active with both ethanol (or acetaldehyde) and steroid substrates, and had similar Michaelis-Menten constants for each of the substrates and coenzymes studied. The three isoenzymes with the lowest migration toward the cathode exhibited the same pH optimum of 10.7 for ethanol oxidation, a greater activity with 5 beta-androstan-3 beta-ol-17-one than with ethanol as a substrate, and an unchanged electrophoretic mobility following storage in the presence of 100 microM dithiothreitol. By contrast the isoenzyme with the highest mobility toward the cathode exhibited a pH optimum of 9.5 for ethanol oxidation, a low steroid/ethanol ratio of activity, and converted to the migrating pattern of the two isoenzymes with intermediate mobility when stored. The similarities between the isoenzymes of rat liver alcohol dehydrogenase differ considerably from differences in substrate specificity exhibited by isoenzymes of horse liver alcohol dehydrogenase.  相似文献   

19.
Cell-free extracts of Hyphomicrobium X showed NAD-dependent aldehyde dehydrogenase activity, provided that NAD addition preceded that of aldehyde. Activity was lost rather rapidly, especially during purification attempts, but this could be partially masked by including a time-dependent restoration step with thiol compounds in the protocol. The nature of the assay buffer appeared to be critical and stimulation occurred on incorporation of K+ ions in the mixture. An even higher specific activity could be achieved by 1,4-dithiothreitol (DTT) treatment of the preparation, followed by removal of DTT, and assaying in the absence of thiol compounds under anaerobic conditions. Exposure of such a preparation to O2 led to a significant decrease in activity within a couple of hours. Immediate inactivation occurred on addition of H2O2, but this could be prevented completely by prior addition of NAD. Since GSH does not participate in the reaction and no stimulating factor was detected, the role of thiol compounds is most probably confined to restoration or prevention of damage to an O2-sensitive, necessary thiol group. Since the same features were found for cell-free extract as for the partially purified enzyme, only one enzyme type seems to be present. Although the enzyme is a general aldehyde dehydrogenase, the kinetic parameters and the specific activity of the cell-free extract for formaldehyde indicate that it may play a role in formaldehyde dissimilation by Hyphomicrobium X. The NAD-linked, GSH- and factor-independent aldehyde dehydrogenase described here appears to be different in several respects from the formaldehyde dehydrogenase of Pseudomonas putida (EC 1.2.1.46) (despite showing similar behavior toward coenzymes and factors) but resembles the aldehyde dehydrogenase from baker's yeast (EC 1.2.1.5).  相似文献   

20.
S M Lau  P Powell  H Buettner  S Ghisla  C Thorpe 《Biochemistry》1986,25(15):4184-4189
The flavoprotein medium-chain acyl coenzyme A (acyl-CoA) dehydrogenase from pig kidney exhibits an intrinsic hydratase activity toward crotonyl-CoA yielding L-3-hydroxybutyryl-CoA. The maximal turnover number of about 0.5 min-1 is 500-1000-fold slower than the dehydrogenation of butyryl-CoA using electron-transferring flavoprotein as terminal acceptor. trans-2-Octenoyl- and trans-2-hexadecenoyl-CoA are not hydrated significantly. Hydration is not due to contamination with the short-chain enoyl-CoA hydratase crotonase. Several lines of evidence suggest that hydration and dehydrogenation reactions probably utilize the same active site. These two activities are coordinately inhibited by 2-octynoyl-CoA and (methylenecyclopropyl)acetyl-CoA [whose targets are the protein and flavin adenine dinucleotide (FAD) moieties of the dehydrogenase, respectively]. The hydration of crotonyl-CoA is severely inhibited by octanoyl-CoA, a good substrate of the dehydrogenase. The apoenzyme is inactive as a hydratase but recovers activity on the addition of FAD. Compared with the hydratase activity of the native enzyme, the 8-fluoro-FAD enzyme exhibits a roughly 2-fold increased activity, whereas the 5-deaza-FAD dehydrogenase is only 20% as active. A mechanism for this unanticipated secondary activity of the acyl-CoA dehydrogenase is suggested.  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