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.
The enzymes tyrosinase, catecholoxidase and hemocyanin all share similar active sites, although their physiological functions differ. Hemocyanins serve as oxygen carrier proteins, and tyrosinases and catecholoxidases (commonly referred to as phenoloxidases in arthropods) catalyze the hydroxylation of monophenols or the oxidation of o-diphenols to o-quinones, or both. Tyrosinases are activated in vivo by limited proteolytic cleavage, which might open up substrate access to the catalytic site. It has recently been demonstrated that if hemocyanins are subjected to similar proteolytic treatments (in vitro) they also exhibit at least catecholoxidase reactivity. On the basis of their molecular structures, hemocyanins are used as model systems to understand the substrate-active-site interaction between catecholoxidases and tyrosinases.  相似文献   

2.
Nillius D  Jaenicke E  Decker H 《FEBS letters》2008,582(5):749-754
Phenoloxidases and hemocyanins have similar type 3 copper centers although they perform different functions. Hemocyanins are oxygen carriers, while phenoloxidases (tyrosinase/catecholoxidase) catalyze the initial step in melanin synthesis. Tyrosinases catalyze two subsequent reactions, whereas catecholoxidases catalyze only the second one. Recent results indicate that hemocyanins can also function as phenoloxidases and here we show for the first time that hemocyanin can be converted to phenoloxidase. Furthermore, its substrate specificity can be switched between catecholoxidase and tyrosinase activity depending on effectors such as hydroxymethyl-aminomethan (Tris) and Mg(2+)-ions. This demonstrates that substrate specificity is not caused by a chemical modification of the active site.  相似文献   

3.
The activity of tyrosinase, the rate-limiting enzyme for melanin synthesis, is higher in Black skin melanocytes than in melanocytes derived from Caucasian skin. This variation in enzyme activity is not due to differences in tyrosinase abundance or tyrosinase gene activity, but, rather, is due to differences in the catalytic activity of preexisting tyrosinase. In melanocytes, tyrosinase is localized to the membrane of melanosomes and in Caucasian melanocytes the melanosome-bound enzyme is largely inactive. Conversely, in melanosomes of Black melanocytes, tyrosinase has high catalytic activity. Treatment of Caucasian melanocytes with the lysosomotropic compound ammonium chloride or with the ionophores nigericin and monensin results in a rapid and pronounced increase in tyrosinase activity. This increase occurs without any change in tyrosinase abundance, indicating that these compounds are increasing the catalytic activity of preexisting enzyme. Inhibition of the vacuolar proton pump V-ATPase by treatment of Caucasian melanocytes with bafilomycin also increases tyrosinase activity. In contrast to the 10-fold increase in tyrosinase observed in Caucasian melanocytes, neither ammonium chloride, monensin, nigericin, nor bafilomycin is able to increase the already high level of tyrosinase activity present in melanosomes of melanocytes derived from Black skin. Finally, staining of Caucasian melanocytes with the fluorescent weak base acridine orange shows that melanosomes of Caucasian, but not Black, melanocytes are acidic organelles. These data support a model for racial pigmentation that is based on differences in melanosome pH in Black and Caucasian skin types. The models suggests that melanosomes of Caucasian melanocytes are acidic, while those of Black individuals are more neutral. Since tyrosinase is inactive in an acid environment, the enzyme is largely inactive in Caucasian melanosomes but fully active in Black melanosomes.  相似文献   

4.
We studied the role of chromatin accessibility and methylation in the banding patterns produced by means of in situ nick-translation (NT) and restriction enzyme (RE) banding techniques. For these studies we used the X chromosomes of Microtus cabrerae because of their large segment with four different types of constitutive heterochromatin and because in these chromosomes we can also compare active and inactive euchromatin. The results demonstrate that constitutive heterochromatin in the X chromosomes of M. cabrerae is methylated at specific sequences in both active and inactive Xs. They also show that NT-based techniques are suitable for detecting weak differences in chromatin accessibility, such as differences between active and inactive euchromatin, and are able to distinguish methylation only at the accessible sites. Thus, when methylation has to be mapped in situ, additional experiments have to be performed in order to distinguish findings due to differential accessibility. RE banding seems less sensitive to slight differences in chromatin accessibility, and might thus be more suitable than in situ NT-based techniques for methylation mapping. In harmony with these results, HpaII-based RE banding is able to distinguish between active and inactive euchromatin, possibly depending on its methylation status.  相似文献   

5.
Several enzymic and physical properties of Sepharose-bound trypsin and activated Sepharose-bound trypsinogen have been compared to those of the soluble enzyme. Sepharose-bound trypsinogen could be activated to the same extent as soluble trypsinogen; the release of the activation peptide and formation of the active site occurred as expected in the presence of catalytic amounts of trypsin. With synthetic substrates, the relative activity and pH dependence of both immobilized trypsin preparations were essentially identical and nearly the same as the soluble enzyme. Sepharose-trypsin also formed an inactive complex with soybean trypsin inhibitor, with 85% of the active sites participating. In contrast, the activity of Sepharose-trypsin with chymotrypsinogen and with trypsinogen as substrates was only 40% that of soluble trypsin. There is evidence for some catalytic heterogeneity of active sites of bound trypsin; probably those sites buried within the gel have a limited catalytic efficiency with macromolecular substrates. The immobilized enzyme is more stable than the soluble enzyme at elevated temperatures and to concentrated urea, and denaturation by urea at pH 8 is fully reversible since the loss of molecules by autolysis is eliminated.  相似文献   

6.
We studied the role of chromatin accessibility and methylation in the banding patterns produced by means of in situ nick-translation (NT) and restriction enzyme (RE) banding techniques. For these studies we used the X chromosomes of Microtus cabrerae because of their large segment with four different types of constitutive heterochromatin and because in these chromosomes we can also compare active and inactive euchromatin. The results demonstrate that constitutive heterochromatin in the X chromosomes of M. cabrerae is methylated at specific sequences in both active and inactive Xs. They also show that NT-based techniques are suitable for detecting weak differences in chromatin accessibility, such as differences between active and inactive euchromatin, and are able to distinguish methylation only at the accessible sites. Thus, when methylation has to be mapped in situ, additional experiments have to be performed in order to distinguish findings due to differential accessibility. RE banding seems less sensitive to slight differences in chromatin accessibility, and might thus be more suitable than in situ NT-based techniques for methylation mapping. In harmony with these results, HpaII-based RE banding is able to distinguish between active and inactive euchromatin, possibly depending on its methylation status.  相似文献   

7.
The structure of tyrosinase (Tyr) is reviewed from a double point of view. On the one hand, by comparison of all Tyr found throughout nature, from prokaryotic organisms to mammals and on the other, by comparison with the tyrosinase related proteins (Tyrps) that appeared late in evolution, and are only found in higher animals. Their structures are reviewed as a whole rather than focused on the histidine (His)-bound metal active site, which is the part of the molecule common to all these proteins. The availability of crystallographic data of hemocyanins and recently of sweet potato catechol oxidase has improved the model of the three-dimensional structure of the Tyr family. Accordingly, Tyr has a higher structural disorder than hemocyanins, particularly at the CuA site. The active site seems to be characterized by the formation of a hydrophobic pocket with a number of conserved aromatic residues sited close to the well-known His. Other regions specific of the mammalian enzymes, such as the cytosolic C-terminal tail, the cysteine clusters, and the N-glycosylation sequons, are also discussed. The complete understanding of the Tyr copper-binding domain and the characterization of the residues determinant of the relative substrate affinities of the Tyrps will improve the design of targeted mutagenesis experiments to understand the different catalytic capabilities of Tyr and Tyrps. This may assist future aims, from the design of more efficient bacterial Tyr for biotechnological applications to the design of inhibitors of undesirable fruit browning in vegetables or of color skin modulators in animals.  相似文献   

8.
Mutational mapping of the catalytic activities of human tyrosinase.   总被引:7,自引:0,他引:7  
Tyrosinase (EC 1.14.18.1) is a copper-containing metalloglycoprotein that catalyzes several steps in the melanin pigment biosynthetic pathway; the hydroxylation of tyrosine to L-3,4-dihydroxyphenylalanine (dopa) and the subsequent oxidation of dopa to dopaquinone. It has been proposed that tyrosinase is also able to oxidize 5,6-dihydroxyindole (DHI), a later product in the melanogenic pathway, to indole-5,6-quinone. Tyrosinase enzymatic activity is deficient in patients with classic type I oculocutaneous albinism (OCA), and more than 50 distinct mutations have now been identified in the tyrosinase genes of such patients. To determine the effects of the various tyrosinase gene mutations on the catalytic activities of the enzyme, we carried out site-directed mutagenesis of human tyrosinase cDNA, transiently expressed the mutant cDNAs in transfected HeLa cells, and assayed the resultant encoded proteins for tyrosine hydroxylase, dopa, and DHI oxidase activities, and resulting melanin production. The tyrosine hydroxylase activity of normal tyrosinase is thermostable, whereas its dopa oxidase and DHI oxidase activities are temperature-sensitive. Although all amino acid substitutions tested generally affected the dopa oxidase and DHI oxidase activities in parallel, several exerted distinctly different effects on the tyrosine hydroxylase activities. Together, these results confirm the DHI oxidase activity of mammalian tyrosinase and suggest that the dopa oxidase and DHI oxidase activities of tyrosinase share a common catalytic site, whereas the tyrosine hydroxylase catalytic site is at least partially distinct in the tyrosinase polypeptide.  相似文献   

9.
The structure of tyrosinase (Tyr) is reviewed from a double point of view. On the one hand, by comparison of all Tyr found throughout nature, from prokaryotic organisms to mammals and on the other, by comparison with the tyrosinase related proteins (Tyrps) that appeared late in evolution, and are only found in higher animals. Their structures are reviewed as a whole rather than focused on the histidine (His)‐bound metal active site, which is the part of the molecule common to all these proteins. The availability of crystallographic data of hemocyanins and recently of sweet potato catechol oxidase has improved the model of the three‐dimensional structure of the Tyr family. Accordingly, Tyr has a higher structural disorder than hemocyanins, particularly at the CuA site. The active site seems to be characterized by the formation of a hydrophobic pocket with a number of conserved aromatic residues sited close to the well‐known His. Other regions specific of the mammalian enzymes, such as the cytosolic C‐terminal tail, the cysteine clusters, and the N‐glycosylation sequons, are also discussed. The complete understanding of the Tyr copper‐binding domain and the characterization of the residues determinant of the relative substrate affinities of the Tyrps will improve the design of targeted mutagenesis experiments to understand the different catalytic capabilities of Tyr and Tyrps. This may assist future aims, from the design of more efficient bacterial Tyr for biotechnological applications to the design of inhibitors of undesirable fruit browning in vegetables or of color skin modulators in animals.  相似文献   

10.
Staphylococcal aureus epidermolytic toxins (ETs) A and B are responsible for the induction of staphylococcal scalded skin syndrome, a disease of neonates and young children. The clinical features of this syndrome vary from localized blisters to severe exfoliation affecting most of the body surface. Comparison of the crystal structures of two subtypes of ETs-rETA (at 2.0 A resolution), rETB (at 2.8 A resolution), and an active site variant of rETA, Ser195Ala at 2.0 A resolution has demonstrated that their overall topology resembles that of a "trypsin-like" serine protease, but with significant differences at the N- and C-termini and loop regions. The details of the catalytic site in both ET structures are very similar to those in glutamate-specific serine proteases, suggesting a common catalytic mechanism. However, the "oxyanion hole," which is part of the catalytic sites of glutamate specific serine proteases, is in the closed or inactive conformation for rETA, yet in the open or active conformation for rETB. The ETs contain a unique amphipathic helix at the N-terminus, and it appears to be involved in optimizing the conformation of the catalytic site residues. Determination of the structure of the rETA catalytic site variant, Ser195Ala, showed no significant perturbation at the active site, establishing that the loss of biological and esterolytic activity can be attributed solely to disruption of the catalytic serine residue. Finally, the crystal structure of ETs, together with biochemical data and mutagenesis studies, strongly confirms the classification of these molecules as "serine proteases" rather than "superantigens."  相似文献   

11.
Incubation of human plasma prorenin (PR), the enzymatically inactive precursor of renin (EC 3.4.23.15), with a number of nonpeptide high-affinity active site-directed renin inhibitors induces a conformational change in PR, which was detected by a monoclonal antibody that reacts with active renin but not with native inactive PR. This conformational change also occurred when inactive PR was activated during exposure to low pH. Nonproteolytically acid-activated PR, and inhibitor-"activated" PR, as well as native PR, were retained on a blue Sepharose column, in contrast to proteolytically activated PR. Kinetic analysis of the activation of plasma prorenin by renin inhibitor (INH) indicated that native plasma contains an open intermediary form of prorenin, PRoi, in which the active site is exposed and which is in rapid equilibrium with the inactive closed form, PRc. PRoi reacts with inhibitor to form a reversible complex, PRoi.INH, which undergoes a conformational change resulting in a tight complex of a modified open form of prorenin, PRo, and the inhibitor, PRoi.INH-->PRo.INH. The PRoi-to-PRo conversion leads to the expression of an epitope on the renin part of the molecule that is recognized by a renin-specific monoclonal antibody. Presumably, PRo corresponds to the enzymatically active form of PR that is formed during exposure to low pH. Thus, it seems that the propeptide of PR interacts with the renin part of the molecule not only at or near the enzyme's active site but also at some distance from the active site. Interference with the first interaction by renin inhibitor leads to destabilization of the propeptide, by which the second interaction is disrupted and the enzyme assumes its active conformation. The results of this study may provide a model for substrate-mediated prorenin activation and increase the likelihood that enzymatically active prorenin is formed in vivo.  相似文献   

12.
We have established by radioimmunoprecipitation that tyrosine-DOPA oxidase (TDO, tyrosinase) [EC 1.14.18.1] is first synthesized by frog embryos at the early neurula stage soon after embryonic induction of the neural plate by the underlying chordamesoderm. The DOPA moiety of the enzyme, at the time of its first appearance, is almost inactive enzymatically and can be activated by mild proteolysis (with trypsin). A very large increase in the amount of active DOPA oxidizing enzyme (without trypsinization) is observed at hatching (stage 21), and this is accompanied by melanin deposition in pigment cells. The tyrosine moiety of the enzyme is also partially inactive at the time of first synthesis, but the ratio of active to inactive enzyme remains approximately constant throughout early development. DOPA decarboxylase enzymatic activity is first detected at neurula stage, and this activity is accompanied by the first appearance of catechol amines.  相似文献   

13.
Gibson LM  Lovelace LL  Lebioda L 《Biochemistry》2008,47(16):4636-4643
Loop 181-197 of human thymidylate synthase (hTS) populates two conformational states. In the first state, Cys195, a residue crucial for catalytic activity, is in the active site (active conformer); in the other conformation, it is about 10 A away, outside the active site (inactive conformer). We have designed and expressed an hTS variant, R163K, in which the inactive conformation is destabilized. The activity of this mutant is 33% higher than that of wt hTS, suggesting that at least one-third of hTS populates the inactive conformer. Crystal structures of R163K in two different crystal forms, with six and two subunits per asymmetric part of the unit cells, have been determined. All subunits of this mutant are in the active conformation while wt hTS crystallizes as the inactive conformer in similar mother liquors. The structures show differences in the environment of catalytic Cys195, which correlate with Cys195 thiol reactivity, as judged by its oxidation state. Calculations show that the molecular electrostatic potential at Cys195 differs between the subunits of the dimer. One of the dimers is asymmetric with a phosphate ion bound in only one of the subunits. In the absence of the phosphate ion, that is in the inhibitor-free enzyme, the tip of loop 47-53 is about 11 A away from the active site.  相似文献   

14.
BphC derived from Pseudomonas sp. strain KKS102, an extradiol type catecholic dioxygenase, is a non-heam iron-containing enzyme, playing an important role in the degradation of biphenyl/PCB (Poly Chlorinated Biphenyls) in the microbe. Although we had earlier solved the crystal structure of KKS102 BphC, it was the inactive form with Fe(III) in the active site. In order to determine the active form structure, BphC was re-activated by anaerobic incubation with Fe(II) and ascorbate, and crystallized anaerobically. The crystal structures of activated BphC and its substrate complex (E x S complex) were determined at 2.0 A resolution under cryogenic condition. In addition, crystal structures of unactivated BphC in substrate free and complex forms were also re-determined. Comparison of activated and unactivated E x S complexes reveals that the orientation of the bound substrate in the active site is significantly different between the two. The structural comparison of the substrate free and complex forms of activated BphC show certain small conformational shifts around the active site upon substrate binding. As a result of the conformational shifts, His194, which has been suggested as the catalytic base, takes part in a weak hydrogen bond with hydroxyl group of the substrate.  相似文献   

15.
BfiI is a novel type IIs restriction endonuclease that, unlike all other restriction enzymes characterised to date, cleaves DNA in the absence of Mg(2+). The amino acid sequence of the N-terminal part of BfiI has some similarities to Nuc of Salmonella typhimurium, an EDTA-resistant nuclease akin to phospholipase D. The dimeric form of Nuc contains a single active site composed of residues from both subunits. To examine the roles of the amino acid residues of BfiI that align with the catalytic residues in Nuc, a set of alanine replacement mutants was generated by site-directed mutagenesis. The mutationally altered forms of BfiI were all catalytically inactive but were still able to bind DNA specifically. The active site of BfiI is thus likely to be similar to that of Nuc. BfiI was also found by gel-filtration to be a dimer in solution. Both gel-shift and pull-down assays indicated that the dimeric form of BfiI binds two copies of its recognition sequence. In reactions on plasmids with either one or two copies of its recognition sequence, BfiI cleaved the DNA with two sites more rapidly than that with one site. Yet, when bound to two copies of its recognition sequence, the BfiI dimer cleaved only one phosphodiester bond at a time. The dimer thus seems to contain two DNA-binding domains but only one active site.  相似文献   

16.
Kinetic experiments are reported showing that mammalian tyrosinase from B16 mouse melanoma is significantly activated by catalytic amounts of ferrous ions. Monitoring of tyrosine oxidation by both dopachrome formation and oxygen consumption showed that ferrous ions at micromolar concentrations induce a marked enzymatic activity with 0.01 U/ml of highly purified tyrosinase, whereas no detectable reaction occurs in the absence of metal over a sufficiently prolonged period of time. The extent of the activating effect, which is specific for the reduced form of iron, is proportional to the concentration of the added metal with a typical saturation profile, no further effect being observed beyond a threshold value. Changing the buffer system from phosphate to hepes or tris results in a marked decrease of the Fe2(+)-induced activation. Scavengers of active oxygen species, such as superoxide dismutase, catalase, formate and mannitol have no detectable effect on the tyrosinase activity. These results are accounted for in terms of an activation mechanism involving reduction of the cupric ions at the active site of the resting enzyme.  相似文献   

17.
The midgut of the yellow mealworm, Tenebrio molitor L. (Coleoptera: Tenebrionidae) larvae has four beta-glycosidases. The properties of two of these enzymes (betaGly1 and betaGly2) have been described elsewhere. In this paper, the characterization of the other two glycosidases (betaGly3 and betaGly4) is described. BetaGly3 has one active site, hydrolyzes disaccharides, cellodextrins, synthetic substrates and beta-glucosides produced by plants. The enzyme is inhibited by amygdalin, cellotriose, cellotetraose and cellopentaose in high concentrations, probably due to transglycosylation. betaGly3 hydrolyzes beta 1,4-glycosidic linkages with a catalytic rate independent of the substrate polymerization degree (k(int)) of 11.9 s(-1). Its active site is formed by four subsites, where subsites +1 and -1 bind glucose residues with higher affinity than subsite +2. The main role of betaGly3 seems to be disaccharide hydrolysis. BetaGly4 is a beta-galactosidase, since it has highest activity against beta-galactosides. It can also hydrolyze fucosides, but not glucosides, and has Triton X-100 as a non-essential activator (K(a)=15 microM, pH 4.5). betaGly4 has two active sites that can hydrolyze p-nitrophenyl beta-galactoside (NPbetaGal). The one hydrolyzing NPbetaGal with more efficiency is also active against methylumbellipheryl beta-D-galactoside and lactose. The other active site hydrolyzes NPbetaFucoside and binds NPbetaGal weakly. BetaGly4 hydrolyzes hydrophobic substrates with high catalytical efficiency and is able to bind octyl-beta-thiogalactoside in its active site with high affinity. The betaGly4 physiological role is supposed to be the hydrolysis of galactolipids that are found in membranes from vegetal tissues. As the enzyme has a hydrophobic site where Triton X-100 can bind, it might be activated by membrane lipids, thus becoming fully active only at the surface of cell membranes.  相似文献   

18.

Background

o-Aminophenols have been long recognised as tyrosinase substrates. However their exact mode of interaction with the enzyme's active site is unclear. Properly vic-substituted o-aminophenols could help gain some insight into tyrosinase catalytic mechanism.

Methods

Eight vic-substituted o-aminophenols belonging to two isomeric series were systematically evaluated as tyrosinase substrates and/or activators and/or inhibitors, by means of spectrophotometric techniques and HPLC-MS analysis. Some relevant kinetic parameters have also been obtained.

Results

Four o-aminophenolic compounds derived from 3-hydroxyorthanilic acid (2-amino-3-hydroxybenzenesulfonic acid) and their four counterparts derived from the isomeric 2-hydroxymetanilic acid (3-amino-2-hydroxybenzenesulfonic acid) were synthesised and tested as putative substrates for mushroom tyrosinase. While the hydroxyorthanilic derivatives were quite inactive as both substrates and inhibitors, the hydroxymetanilic compounds on the contrary all acted as substrates for the enzyme, which oxidised them to the corresponding phenoxazinone derivatives.

General significance

Based on the available structures of the active sites of tyrosinases, the different affinities of the four metanilic derivatives for the enzyme, and their oxidation rates, we propose a new hypothesis regarding the interaction between o-aminophenols and the active site of tyrosinase that is in agreement with the obtained experimental results.  相似文献   

19.
The catalytic domain of a mosquitocidal toxin prolonged by a C-terminal 44 residue linker connecting to four ricin B-like domains was crystallized. Three crystal structures were established at resolutions between 2.5A and 3.0A using multi-wavelength and single-wavelength anomalous X-ray diffraction as well as molecular replacement phasing techniques. The chainfold of the toxin fragment corresponds to those of ADP-ribosylating enzymes. At pH 4.3 the fragment is associated in a C(7)-symmetric heptamer in agreement with an aggregate of similar size observed by size-exclusion chromatography. In two distinct crystal forms, the heptamers formed nearly spherical, D(7)-symmetric tetradecamers. Another crystal form obtained at pH 6.3 contained a recurring C(2)-symmetric tetramer, which, however, was not stable in solution. On the basis of the common chainfold and NAD(+)-binding site of all ADP-ribosyl transferases, the NAD(+)-binding site of the toxin was assigned at a high confidence level. In all three crystal forms the NAD(+) site was occupied by part of the 44 residue linker, explaining the known inhibitory effect of this polypeptide region. The structure showed that the cleavage site for toxin activation is in a highly mobile loop that is exposed in the monomer. Since it contains the inhibitory linker as a crucial part of the association contact, the observed heptamer is inactive. Moreover, the heptamer cannot be activated by proteolysis because the activation loop is at the ring center and not accessible for proteases. Therefore the heptamer, or possibly the tetradecamer, seems to represent an inactive storage form of the toxin.  相似文献   

20.
Inactivation of the isolated ATPase portion of ATP synthase from beef-heart mitochondria (F1) by its natural inhibitor protein (IP) during steady-state ATP hydrolysis is accompanied by a trapping of 1 mol nucleotide/mol F1 in one of the catalytic sites. The trapped nucleotide is not released during incubation of IP-inhibited F1 in the presence of MgATP at pH 8.0 for at least 20 min, indicating a very low turnover rate of the IP.F1 complex. The ATP/ADP ratio of the trapped nucleotides is higher than that found for transitorily bound nucleotides under the same conditions but in the absence of IP. The IP impairs the acceleration of ATP hydrolysis and product release steps that results from the binding of ATP to an alternate catalytic site. It also inhibits ATP hydrolysis by a single catalytic site or shifts the equilibrium toward ATP formation from bound ADP and Pi. At high pH, an active acidic form of the free IP is transformed to the inactive basic one with a half-time of 3-4 s. This process seems to be prevented by IP binding to F1. The inactive basic form of IP does not compete with the active acidic IP for the binding to F1. The data do not favor the existence of a long-lived catalytically active IP.F1 intermediate during IP action on F1. The reactivation of IP-inhibited membrane-bound F1 by energization may be due to a conformational change in the IP.F1 complex allowing the transformation of IP into an inactive basic state that rapidly dissociates.  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