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.
The action of pig pepsin on a variety of small peptides including Leu-Trp-Met-Arg, Leu-Trp-Met, Leu-Leu-NH2, benzyloxycarbonyl-Phe-Leu and Gly-Leu-Tyr was studied. Leu-Leu-Leu was found to be the major product from the substrates Leu-Trp-Met-Arg and Leu-Trp-Met, indicating that the predominant reaction at pH 3.4 was a transpeptidation of the acyl-transfer type. Leu-Leu-Leu was also formed in high yield by amino transfer from benzyloxycarbonyl-Phe-Leu. Like the amino-transfer reactions the acyl transfer proceeded via a covalent intermediate, since [14C]leucine was not incorporated into transpeptidation products and did not exchange with enzyme-bound leucine in the presence of acceptors. With Leu-Trp-Met both acyl and amino transpeptidation products, namely Leu-Leu, Leu-Leu-Leu, Met-Met and Met-Met-Met, were formed in addition to methionine and leucine. With Leu-Trp-Met-Arg (1 mM) the pH optimum for the rates of hydrolysis and acyl transfer is about pH 3.4. At this pH the rate of acyl transfer exceeds that of hydrolysis; at pH 2, however, hydrolysis was faster than transfer. A comparison of the effect of the length of substrates and products on the reaction rates allows the conclusion that the binding site can extend over eight to nine amino acid residues. Although the experiments provide no conclusive evidence for or against the involvement of amino and/or acyl intermediates in the hydrolysis of long peptides and proteins, the high yield of transpeptidation reactions of both types observed with some substrates suggests a major role for the intermediates in pepsin-catalysed reactions. The results also show that when pig pepsin is used for the digestion of proteins for sequence work, the likelihood of the formation of transpeptidation products is considerable. In this way peptides not present in the original sequence could easily form in a reasonably good yield.  相似文献   

2.
Penicillopepsin catalyses transpeptidation reactions involving the transfer of the N-terminal amino acids of suitable substrates via covalent acyl intermediates to acceptor peptides, usually the substrate. The major products obtained when Phe-Tyr-Thr-Pro-Lys-Ala and Met-Leu-Gly were used as substrates were Phe-Phe and Met-Met respectively. With Met-Leu-Gly the tetrapeptide Met-Met-Leu-Gly was observed as probable intermediate. Co-incubation of Leu-Tyr-Leu and Phe-Tyr-Thr-Pro-Lys-Ala led to the formation of Leu-Phe and Phe-Leu as well as Leu-Leu and Phe-Phe. No reaction was observed with tripeptides in which the first or second amino acid is glycine. It appears that two amino aicds with large hydrophobic residues are needed for the transpeptidation reaction. Nucleophilic compounds other than peptides, such as hydroxylamine, aliphatic alcohols and dinitrophenylhydrazine, were not acceptors for the acyl group. Leucine, phenylalanine and leucine methyl ester also had no effect on the reaction. The transpeptidation reaction proceeded readily at pH 3.6 and 4.7. At pH 6.0 the reaction was slow and at pH 1.9 little or no transpeptidation was observed. Porcine pepsin catalyses similar transpeptidation reactions. Sequence studies show that porcine pepsin and penicillopepsin are homologous. The present study also suggests that they have a very similar mechanism. Evidence available at this time indicates that the mechanism of these enzymes is complex and may be modulated by secondary substrate-enzyme interactions. A hypothesis is presented which proposes that pepsin-catalysed reactions proceed via different covalent intermediates (amino-intermediates or acylintermediates) depending on the nature of the substrate. The possibility that some reactions do not involve covalent intermediates is also discussed.  相似文献   

3.
Thermolysin showed peculiar transpeptidation reactions. Leu-Leu and/or Leu-Leu-Leu were produced at ca. pH 7 from Leu-Leu-NH2 and Cbz-Leu-Leu. Isotope experiments indicated that the transpeptidation products did not use leucine released from the substrates as an acceptor. With Leu-Trp-Met, Leu-Leu, Leu-Leu-Leu and Met-Met were produced as transpeptidation products. A comparative study was done with α-chymotrypsin and pepsin. These results would indicate that thermolysin catalyzed reactions proceed via both acyl and amino intermediates depending upon the substrates, which has been proposed for the mechanism of pepsin. This may also be true in some cases for chymotrypsin and other proteases, which have been known as enzymes of the acyl-enzyme mechanism.  相似文献   

4.
The kinetics of sheep kidney gamma-glutamyl transpeptidase was studied using a novel substrate L-alpha-methyl-gamma-glutamyl-L-alpha-aminobutyrate. When the substrate was incubated with the enzyme in the presence of an amino acid or peptide acceptor, the corresponding L-alpha-methyl-gamma-glutamyl derivatives of the acceptors were formed. In the absence of acceptor only hydrolysis occurred, and no transpeptidation products were detected. The presence of the methyl group on the alpha-carbon apparently prevents enzymatic transfer of the L-alpha-methyl-gamma-glutamyl residue to the amino group of the substrate itself (autotranspeptidation). When the enzyme was incubated with conventional substrates, such as glutathione or gamma-glutamyl-p-nitroanilide and an amino acid acceptor, hydrolysis, autotranspeptidation, and transpeptidation to the acceptor occurred concurrently. Initial velocity measurements in which the concentration of L-alpha-methyl-gamma-glutamyl-L-alpha-aminobutyrate was varied at several fixed acceptor concentrations, and either the release of alpha-aminobutyrate or the formation of the transpeptidation products was determined, yielded results which are consistent with a ping-pong mechanism modified by a hydrolytic shunt. A scheme of such a mechanism is presented. This mechanism predicts the formation of an alpha-methyl-gamma-glutamyl-enzyme intermediate, which can react with an amino acid to form the transpeptidation product; or in the absence of, or in the presence of low concentrations of amino acids, can react with water to form the hydrolytic products. Kinetic derivations for the reaction of the enzyme with the conventional substrate gamma-glutamyl-p-nitroanilide predict either linear or nonlinear double-reciprocal plots, depending on the prevalence of the hydrolytic, autotranspeptidation, or transpeptidation reactions. The results of kinetic experiments confirmed these predictions.  相似文献   

5.
1. The inhibition of pepsin-catalysed hydrolysis of N-acetyl-l-phenylalanyl-l-phenylalanylglycine by products and product analogues was studied. 2. The non-competitive nature of the inhibition by the product N-acetyl-l-phenylalanine confirms an ordered release of products, and points to a common mechanism (involving an amino-enzyme) for pepsin-catalysed transpeptidation and hydrolysis reactions. 3. N-Acetyl-l-phenylalanine ethyl ester is also a non-competitive inhibitor, but here the inhibition is of the ;dead-end' type. No ethanol is detectable in reaction mixtures, indicating that this ester cannot act as an amino group acceptor in a transpeptidation process. 4. The same is true for N-methanesulphonyl-l-phenylalanine methyl and methyl thiol esters. No methanethiol is liberated when the methyl thiol ester is present as an inhibitor of the hydrolytic reaction, and the hope that such a thiol ester would effectively trap the amino-enzyme was not fulfilled.  相似文献   

6.
  • 1.1. The tetrapeptide Ala2-Nph2 (where Nph = p-nitrophenylalanyl) is treated by porcine pepsin to study the mechanism of aminotranspeptidation reactions.
  • 2.2. The major initial product is Ala2-Nph and the major transpeptidation products are Nph2 and Nph3 accompanied by some Nph, a little Nph4, Ala2-Nph3 and Ala2-Nph4.
  • 3.3. Oligomers of Nph greater than tetramers are formed near the end of the reaction.
  • 4.4. In presence of [3H]Nph, no incorporation of Nph into the transpeptidation products is observed.
  • 5.5. 18O-Iabeling shows extensive incorporation of 18O atoms from [18O]water in the carbonyl oxygens of Nph residues.
  相似文献   

7.
The specificity of the Streptomyces R61 penicillin-sensitive D-Ala-D-Ala peptidase has been re-examined with the help of synthetic substrates. The products of the transpeptidation reactions obtained with Gly-L-Xaa dipeptides as acceptor substrates are themselves poor substrates of the enzyme. This is in apparent contradiction with the classically accepted specificity rules for D-Ala-D-Ala peptidases. The Gly-L-Xaa dipeptide is regenerated by both the hydrolysis and transpeptidation reactions. The latter reaction is observed when another Gly-L-Xaa peptide or D-Alanine are supplied as acceptors. Utilization of substrates in which the terminal -COO(-) group has been esterified or amidated shows that a free carboxylate is not an absolute prerequisite for activity. The results are discussed in the context of the expected reversibility of the transpeptidation reaction.  相似文献   

8.
1. The inhibition of pepsin-catalysed hydrolysis of N-acetyl-l-phenylalanyl-l-phenylalanylglycine by products and product analogues was studied. 2. Inhibitors of the l-configuration give rise to linear non-competitive inhibition, whereas those of the d-configuration show linear competitive behaviour. 3. Non-competitive inhibition by the product N-acetyl-l-phenylalanine indicates an ordered release of products, which supports a common mechanism (involving an ;amino-enzyme') for pepsin-catalysed transpeptidation and hydrolysis reactions. 4. The differences in the types of inhibition caused by product analogues of the l- and d-series emphasize the stereospecificity of the binding of these inhibitors to free enzyme and to the putative amino-enzyme intermediate. 5. The results suggest that it is the anion of the acyl product that is released first in the hydrolytic reaction (see Kitson & Knowles, 1971).  相似文献   

9.
The effects of parathyroid hormone (PTH), dihydroxycholecalciferol (1,25-(OH)2 D3), thrombin, epidermal growth factor (EGF) and 12-o-tetradecanoylphorbol-13-acetate (PMA) on the biosynthesis and release of arachidonic acid metabolites were studied in primary cultures of osteoblast-like cells isolated from 18-day-old chick embryo calvaria. Cells were labelled with (14C)-arachidonic acid for 30 h. The radioactive eicosanoids were extracted from the cell culture media after a further 30 h stimulation period and analysed on a PRP-1 column by HPLC. The radioactive products were characterized by co-elution of (3H) standard prostanoids. Osteoblasts showed a basal release of the prostanoids 6-keto-PGF1 alpha, TXB2, PGF2 alpha, PGE2, PGD2 and PGB2, the latter being the most abundant one. Indomethacin (10(-5) M) effectively inhibited the basal release, but not that of an as yet unidentified compound. The release of prostanoids was stimulated by PTH (2 U/ml), thrombin (0.4 NIH/ml), EGF (50 ng/ml) and PMA (25 ng/ml), the latter being by far the most potent one. 1,25-(OH)2D3 was found to slightly inhibit the prostanoid release. These results indicate: (1) primary cultures of osteoblasts synthesize several prostaglandins, thromboxane B2 and one unidentified product. (2) the action on bone of PTH and the various drugs tested may be, at least partly, mediated by an increased prostaglandin production by osteoblasts. Clearly this does not apply to 1,25-(OH)2D3.  相似文献   

10.
Specificity of pepsin-catalyzed peptide bond synthesis   总被引:1,自引:0,他引:1  
The rates of the pepsin-catalyzed synthesis of oligopeptides of the general type A-Phe-Leu-B by the condensation of A-Phe-OH with H-Leu-B have been determined by means of analytical high performance liquid chromatography. Variation of the A group led to large changes in the initial rates of the condensation reaction, and the effect of such changes was found to be similar to that previously found for the secondary specificity of pepsin in the hydrolysis of oligopeptide substrates. Replacement of the Phe and Leu residues of A-Phe-OH or H-Leu-B by other amino acid residues gave relative rates of synthesis in accord with the known primary specificity of the hydrolytic action of pepsin. Partially-acetylated pepsin, which exhibits enhanced hydrolytic activity, also catalyzed the condensation reaction more effectively. The results are discussed in relation to the potential utility and limitations of pepsin as a catalyst in the preparative synthesis of oligopeptides and to the problem of the mechanism of its action.  相似文献   

11.
The high molecular weight of chitosan, which results in a poor solubility at neutral pH values and high viscosity aqueous solutions, limits its potential uses in the fields of food, health and agriculture. However, most of these limitations are overcome by chitosan oligosaccharides obtained by enzymatic hydrolysis of the polymer. Several commercial enzymes with different original specificities were assayed for their ability to hydrolyze a 93% deacetylation degree chitosan and compared with a chitosanase. According to the patterns of viscosity decrease and reducing end formation, three enzymes--cellulase, pepsin and lipase A--were found to be particularly suitable for hydrolyzing chitosan at a level comparable to that achieved by chitosanase. Unlike the appreciable levels of both 2-amino-2-deoxy-D-glucose and 2-acetamido-2-deoxy-D-glucose monomers released from chitosan by the other enzymes after a 20h-hydrolysis (4.6-9.1% of the total product weight), no monomer could be detected following pepsin cleavage. As a result, pepsin produced a higher yield of chitosan oligosaccharides than the other enzymes: 52% versus as much as 46%, respectively. Low molecular weight chitosans accounted for the remaining 48% of hydrolysis products. The calculated average polymerization degree of the products released by pepsin was around 16 units after 20h of hydrolysis. This product pattern and yield are proposed to be related to the bond cleavage specificity of pepsin and the high deacetylation degree of chitosan used as substrate. The optimal reaction conditions for hydrolysis of chitosan by pepsin were 40 degrees C and pH 4.5, and an enzyme/substrate ratio of 1:100 (w/w) for reactions longer than 1h.  相似文献   

12.
Penicillopepsin acting on Nph-Ala2-amide (where Nph = p-nitrophenylalanyl) catalyzes a transpeptidation reaction which leads to the formation of Nph2-Ala2-amide, which arises from condensation of the substrate with enzyme-bound Nph, as the first product released from the enzyme. This is followed by a stage during which Nph3 and Ala2-amide are the major products. A small amount of Nph4 is also formed during this time. Nph and Nph2, formed during the reactions, are tightly, but probably not covalently, bound to the enzyme. They appear as free products only as a result of the cleavage of Nph3 and Nph4 and after most of the substrate Nph-Ala2-amide has been used up. They act as acceptors for the substrate and for Nph2-Ala2-amide. Nph3-Ala2-amide, formed by condensation of Nph-Ala2-amide or of Nph2-Ala2-amide with enzyme-bound Nph2 or Nph, respectively, is also released but is cleaved rapidly to give Nph3 and Ala2-amide. Incorporation of 18O from [18O]water into the carbonyl oxygens of the products is extensive and shows that release of the intermediates is slower than peptide bond cleavage and peptide bond formation. Hence the rate-limiting step in these reactions is product release. No 18O is incorporated into the initial substrate. We propose that Nph and Nph2 as intermediates are held in the active site by hydrogen bonds and by two strong electrostatic interactions.  相似文献   

13.
The antioxidative action of fullerenes has received much attention, but their reaction mechanism toward lipid-derived peroxyl radicals has not been well elucidated. In this study, the reaction products of [60]fullerene (C(60)) during the autoxidation of methyl linoleate (MeL) were isolated and their structures were characterized. MeL containing 0.1mol% C(60) was autoxidized at 60°C in bulk phase and two reaction products of C(60), 1 and 2, were obtained. The structure of 1 was the addition products of C(60) with 9-peroxyl-10-alkyl radicals of methyl (11E)-13-hydroperoxy-11-octadecaenoate (1a and 1b) and with 12-alkyl-13-peroxyl radicals of methyl (10E)-9-hydroperoxy-10-octadecaenoate (1c and 1d). 2 was a mixture of the addition products of C(60) with 9,11-dialkyl radicals of methyl 9,12-octadecadienoate (2a) and with 11,13-dialkyl radicals of methyl 9,12-octadecadienoate (2b). When MeL containing 0.1mol% C(60) was autoxidized at 60°C under air-sufficient and air-insufficient conditions, C(60) could suppress the formation of MeL hydroperoxides in both conditions. The reaction product of C(60) first formed was 2 even under air-sufficient conditions, and then 1 was accumulated. The results indicate that the primary antioxidative action of C(60) would be trapping of chain-initiating carbon-centered radicals of unsaturated lipid to form 2.  相似文献   

14.
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   

15.
When erythrocytes are exposed to hypoxia, hemoglobin (Hb) stabilizes in the T-state by capturing 2,3-bisphosphoglycerate. This process could reduce the intracellular pool of glycolytic substrates, jeopardizing cellular energetics. Recent observations suggest that hypoxia-induced activation of glycolytic enzymes is correlated with their release from Band III (BIII) on the cell membrane. Based on these data, we developed a mathematical model of erythrocyte metabolism and compared hypoxia-induced differences in predicted activities of the enzymes, their products, and cellular energetics between models with and without the interaction of Hb with BIII. The models predicted that the allostery-dependent Hb interaction with BIII accelerates consumption of upstream glycolytic substrates such as glucose 6-phosphate and increases downstream products such as phosphoenolpyruvate. This prediction was consistent with metabolomic data from capillary electrophoresis mass spectrometry. The hypoxia-induced alterations in the metabolites resulted from acceleration of glycolysis, as judged by increased conversion of [(13)C]glucose to [(13)C]lactate. The allostery-dependent interaction of Hb with BIII appeared to contribute not only to maintenance of energy charge but also to further synthesis of 2,3-bisphosphoglycerate, which could help sustain stabilization of T-state Hb during hypoxia. Furthermore, such an activation of glycolysis was not observed when Hb was stabilized in R-state by treating the cells with CO. These results suggest that Hb allostery in erythrocytes serves as an O(2)-sensing trigger that drives glycolytic acceleration to stabilize intracellular energetics and promote the ability to release O(2) from the cells.  相似文献   

16.
Wang C  Ye S  Dai L  Liu X  Tong Z 《Carbohydrate research》2007,342(15):2237-2243
Polyelectrolyte multilayer films were prepared through layer-by-layer (LbL) self-assembly of chitosan (CHI) and pyrene labeled poly(2-acrylamido-2-methylpropanesulfonic acid) (APy). After incubation in an enzyme pepsin solution, multilayer films were partially destroyed as detected by a decrease in fluorescence intensity due to enzymatic degradation of CHI and desorption of APy. The multilayer desorption rate was the highest at pH 4.0. Increasing temperature from 20 degrees C to 60 degrees C accelerated desorption. The enzymatic desorption was also observed from microcapsule walls made of CHI/alginate (ALG) multilayer films directly deposited on indomethacin (IDM) microcrystals by LbL self-assembly. After pepsin erosion, the IDM release from the microcapsule monitored by UV absorbance was obviously accelerated due to desorption. The influence of incubation time, pH, and temperature of the pepsin solution on the IDM release was investigated. The release rate was the fastest after incubation in the pepsin solution at pH 4.0 due to the highest activity of pepsin. Increasing incubation temperature from 20 degrees C to 60 degrees C, however, slowed down the release rate, which was considered to be due to the formation of more perfect and compact multilayer films through the chain rearrangement at higher temperatures. The CHI/ALG multilayer film was found to maintain its barrier function to the IDM diffusion even after 6-h incubation in the pepsin solution.  相似文献   

17.
Degradation of dermorphin, [D-Arg2]dermorphin and [D-Arg2, Gly3, Phe4]dermorphin in a soluble rat brain extract was examined. The former two heptapeptides were degraded in a similar fashion to produce corresponding N-terminal tetrapeptide as the main degradation product along with the parallel release of Tyr5, Pro6 and Ser7-NH2. Tyr-D-Arg-Phe-Gly showed a good enzymatic stability. When captopril, an angiotensin-converting enzyme inhibitor, was present in the incubation mixture, hydrolysis of the Gly4-Tyr5 bond was markedly suppressed and resulted in release of the corresponding N-terminal hexapeptide as the main degradation product. Combined use of captopril and amastatin, an aminopeptidase inhibitor, markedly suppressed the hydrolysis of these peptides. On the other hand, [D-Arg2, Gly3, Phe4]dermorphin was hydrolyzed easier than the other two heptapeptides and considerable amounts of Tyr1 and Phe4 were released after 20 hr incubation while the N-terminal tetrapeptide, Tyr-D-Arg-Gly-Phe, showed a good enzymatic stability. On the basis of these results, possible degradation pathways of these heptapeptides were discussed.  相似文献   

18.
Highly purified, but not homogeneous, samples of helix-destabilizing protein 1 from mouse myeloma contain a novel oligonucleotide-releasing DNA exonuclease. This enzyme was separated from helix-destabilizing protein 1 and obtained in highly purified form. A polypeptide of Mr 41 000 is a main constituent of the purified enzyme, and this polypeptide comigrated with the exonuclease activity during the final step of the purification, Sephacryl S-200 gel filtration where the enzyme had a native Mr of 40 000. Overall purification of enzyme activity was greater than 20 000-fold. This exonuclease releases 5'-oligonucleotides in a limited processive manner in both the 5'----3' and 3'----5' directions. Activity of the enzyme is resistant to 1 mM N-ethylmaleimide, requires a divalent cation, has an alkaline pH optimum, and degrades single-stranded DNA much faster than double-stranded DNA or RNA. The predominant oligonucleotide product with uniformly labeled substrates is (pdN)2. With 3' end labeled substrates, greater than 95% of the labeled products are (pdN)4 and (pdN)5; with 5' end labeled substrates, the main labeled product is (pdA)2. The rate of product release from 3' and 5' end labeled substrates is nearly identical at 37 degrees C. A model of the action of this enzyme and a comparison with a human placenta exonuclease [Doniger, J., & Grossman, L. (1976) J. Biol. Chem. 251, 4579-4587] are discussed.  相似文献   

19.
The proteolytic activities of α-chymotrypsin, trypsin, pepsin, bromelain, and an extract from germinating pumpkin seeds were determined by their ability to effect the release of 1-anilino-8-naphthalenesulfonate bound to internal hydrophobic sites in intact protein substrates resulting in a decline in fluorescence. Casein, glyceraldehyde-3-P dehydrogenase, urease, catalase, pumpkin seed globulin, and bovine serum albumin enhanced the fluorescence of 1-anilino-8-naphthalenesulfonate sufficiently to be used as proteolytic substrates in this assay procedure. The activity of 1 μg chymotrypsin or trypsin and 100 ng pepsin could be easily detected by this method within 4 to 8 min depending upon the protein substrate. The digestive enzymes and bromelain exhibited activity against most if not all six of the protein substrates used. In contrast, the extract from germinating pumpkin seeds exhibited significant activity only against pumpkin seed globulin, with maximal activity at pH 7.4. Compared with the assay method for proteolytic activity utilizing ninhydrin analysis of the reaction products, this method was at least 10 times more rapid and gave significant detectable activity with much lower quantities of proteolytic enzyme.  相似文献   

20.
Ribosome serves as a universal molecular machine capable of synthesis of all the proteins in a cell. Small-molecule inhibitors, such as ribosome-targeting antibiotics, can compromise the catalytic versatility of the ribosome in a context-dependent fashion, preventing transpeptidation only between particular combinations of substrates. Classic peptidyl transferase center inhibitor chloramphenicol (CHL) fails to inhibit transpeptidation reaction when the incoming A site acceptor substrate is glycine, and the molecular basis for this phenomenon is unknown. Here, we present a set of high-resolution X-ray crystal structures that explain why CHL is unable to inhibit peptide bond formation between the incoming glycyl-tRNA and a nascent peptide that otherwise is conducive to the drug action. Our structures reveal that fully accommodated glycine residue can co-exist in the A site with the ribosome-bound CHL. Moreover, binding of CHL to a ribosome complex carrying glycyl-tRNA does not affect the positions of the reacting substrates, leaving the peptide bond formation reaction unperturbed. These data exemplify how small-molecule inhibitors can reshape the A-site amino acid binding pocket rendering it permissive only for specific amino acid residues and rejective for the other substrates extending our detailed understanding of the modes of action of ribosomal antibiotics.  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