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.
Bradyrhizobium japonicum strain 110spc4 was capable of chemolithoautotrophic growth with carbon monoxide (CO) as a sole energy and carbon source under aerobic conditions. The enzyme carbon monoxide dehydrogenase (CODH; EC 1.2.99.2) has been purified 21-fold, with a yield of 16% and a specific activity of 58 nmol of CO oxidized/min/mg of protein, by a procedure that involved differential ultracentrifugation, anion-exchange chromatography, hydrophobic interaction chromatography, and gel filtration. The purified enzyme gave a single protein and activity band on nondenaturing polyacrylamide gel electrophoresis and had a molecular mass of 230,000 Da. The 230-kDa enzyme was composed of large (L; 75-kDa), medium (M; 28.4-kDa), and small (S; 17.2-kDa) subunits occurring in heterohexameric (LMS)(2) subunit composition. The 75-kDa polypeptide exhibited immunological cross-reactivity with the large subunit of the CODH of Oligotropha carboxidovorans. The B. japonicum enzyme contained, per mole, 2.29 atoms of Mo, 7.96 atoms of Fe, 7.60 atoms of labile S, and 1.99 mol of flavin. Treatment of the enzyme with iodoacetamide yielded di(carboxamidomethyl)molybdopterin cytosine dinucleotide, identifying molybdopterin cytosine dinucleotide as the organic portion of the B. japonicum CODH molybdenum cofactor. The absorption spectrum of the purified enzyme was characteristic of a molybdenum-containing iron-sulfur flavoprotein.  相似文献   

2.
CO dehydrogenase (EC 1.2.99.2) catalyzes the oxidation of CO according to the following equation: CO + H2O-->CO2 + 2 e- + 2 H+. It is a selenium-containing molybdo-iron-sulfur-flavoenzyme, which has been crystallized and structurally characterized in its oxidized state from the aerobic CO utilizing bacteria Oligotropha carboxidovorans and Hydrogenophaga pseudoflava. Both CO dehydrogenase structures show only minor differences, and the enzymes are dimers of two heterotrimers. Each heterotrimer is composed of a molybdoprotein, a flavoprotein, and an iron-sulfur protein. CO oxidation takes place at the molybdoprotein which contains a 1:1 mononuclear complex of molybdopterin-cytosine dinucleotide and a Mo-ion, along with a catalytically essential S-selanylcysteine. The latter is appropriately positioned in the SeMo-active site by a unique VAYRCSFR active site loop. In H. pseudoflava the arginine preceeding the cysteine in the active site loop is modified to a Cgamma-hydroxy arginine residue which has no obvious function. The substituents in the first coordination sphere of the Mo-ion are the enedithiolate sulfur atoms of the molybdopterin-cytosine dinucleotide, two oxo- and a sulfido-group. Extended X-ray absorption fine structure spectroscopy (EXAFS), along with the crystal structure of CO dehydrogenase (23.2 U mg(-1)) at 1.85 A resolution, have identified a sulfur atom at 2.3 A from the Mo-ion. The sulfur reacts with cyanide yielding thiocyanate. The corresponding inactive desulfo-CO dehydrogenase shows a typical desulfo inhibited-type of Mo-electron paramagnetic resonance (EPR) spectrum. Structural changes at the SeMo-site during catalysis are suggested by the Mo to Se distance of 3.7 A and the Mo-S-Se angle of 113 degrees in the oxidized enzyme which increase to 4.1 A, and 121 degrees, respectively, in the reduced enzyme. The intramolecular electron transport chain in CO dehydrogenase involves the following prosthetic groups and minimal distances: CO-->[Mo of the molybdenum cofactor] - 14.6 A - [2Fe-2S]I - 12.4 A - [2Fe-2S]II - 8.7 A - [FAD].  相似文献   

3.
The transition element molybdenum (Mo) is an essential micronutrient for plants where it is needed as a catalytically active metal during enzyme catalysis. Four plant enzymes depend on molybdenum: nitrate reductase, sulphite oxidase, xanthine dehydrogenase, and aldehyde oxidase. However, in order to gain biological activity and fulfil its function in enzymes, molybdenum has to be complexed by a pterin compound thus forming the molybdenum cofactor. In this article, the path of molybdenum from its uptake into the cell, via formation of the molybdenum cofactor and its storage, to the final modification of the molybdenum cofactor and its insertion into apo-metalloenzymes will be reviewed.  相似文献   

4.
The structurally characterized molybdoenzyme carbon monoxide dehydrogenase (CODH) catalyzes the oxidation of CO to CO2 in the aerobic bacterium Oligotropha carboxidovorans. The active site of the enzyme was studied by Mo- and Cu-K-edge X-ray absorption spectroscopy. This revealed a bimetallic [Cu(I)SMo(VI)(double bond O)2] cluster in oxidized CODH which was converted into a [Cu(I)SMo(IV)(double bond O)OH2] cluster upon reduction. The Cu...Mo distance is 3.70 A in the oxidized form and is increased to 4.23 A upon reduction. The bacteria contain CODH species with the complete and functional bimetallic cluster along with enzyme species deficient in Cu and/or bridging S. The latter are precursors in the posttranslational biosynthesis of the metal cluster. Cu-deficient CODH is the most prominent precursor and contains a [HSMo(double bond O)OH2] cluster. Se-K-edge X-ray absorption spectroscopy demonstrates that Se is coordinated by two C atoms at 1.94-1.95 A distance. This is interpreted as a replacement of the S in methionine residues. In contrast to a previous report [Dobbek, H., Gremer, L., Meyer, O., and Huber, R. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 8884-8889] Se was not identified in the active site of CODH.  相似文献   

5.
Xanthine dehydrogenase has been purified from Pseudomonas aeruginosa cultured on a rich medium and induced with hypoxanthine. The enzyme was shown to contain FAD, iron sulfur centers and a molybdenum cofactor as prosthetic groups. Analysis of the molybdenum cofactor in this enzyme has revealed that the cofactor contains molybdopterin (MPT) rather than molybdopterin guanine dinucleotide or molybdopterin cytosine dinucleotide which have previously been identified in a number of molybdoenzymes of bacterial origin. The pterin cofactor in P.aeruginosa xanthine dehydrogenase was alkylated and the resulting product was identified as dicarboxamidomethyl molybdopterin. In addition, the pterin released from the enzyme by denaturation with guanidine-HCl was found to chromatograph on Sephadex G-15 with an apparent molecular weight of 350. These results document the first example of a bacterial enzyme with a molybdenum cofactor comprising molybdopterin and the metal only.  相似文献   

6.
In dimethylsulfoxide reductase of Rhodobacter capsulatus tryptophan-116 forms a hydrogen bond with a single oxo ligand bound to the molybdenum ion. Mutation of this residue to phenylalanine affected the UV/visible spectrum of the purified Mo(VI) form of dimethylsulfoxide reductase resulting in the loss of the characteristic transition at 720 nm. Results of steady-state kinetic analysis and electrochemical studies suggest that tryptophan 116 plays a critical role in stabilizing the hexacoordinate monooxo Mo(VI) form of the enzyme and prevents the formation of a dioxo pentacoordinate Mo(VI) species, generated as a consequence of the dissociation of one of the dithiolene ligands of the molybdopterin cofactor from the Mo ion.  相似文献   

7.
Isoquinoline 1-oxidoreductase (Ior) from Brevundimonas diminuta 7, encoded by iorAB, is a molybdenum hydroxylase containing a molybdopterin cytosine dinucleotide molybdenum cofactor (Mo-MCD) and two distinct [2Fe2S] clusters. The iorAB genes were inserted into pJB653, generating pIL1. Pseudomonas putida KT2440, and P. putida 86 which produces a Mo-MCD-containing quinoline 2-oxidoreductase when grown on quinoline, were used as recipients for pIL1. Upon induction of gene expression, both clones produced Ior protein, but Ior activity was not detectable in P. putida KT2440 pIL1. In P. putida 86 pIL1, formation of catalytically active Ior required the presence of quinoline, suggesting that accessory gene(s) encoding product(s) essential for the assembly of catalytically competent Ior is (are) part of the quinoline regulon in P. putida 86.  相似文献   

8.
Abstract Respiratory nitrate reductase from the denitrifying bacterium Pseudomonas stutzeri is an iron-sulfur enzyme containing the molybdenum cofactor. Hydrolysis of native nitrate reductase with aqueous sulfuric acid revealed 0.92 mol of 5'-GMP per mol of enzyme. The pterin present in the molybdenum cofactor was liberated from the protein and reacted with iodoacetamide. The resulting di(carboxamidomethyl) (cam) derivative was purified on a C18-cartridge and analyzed for its structural elements. Treatment of the cam derivative with nucleotide pyrophosphatase and subsequent HPLC analysis revealed the formation of di(cam)molybdopterin and 5'-GMP at a 1:1 molar ratio and with a yield of 79% with respect to the molybdenum content of the enzyme. Treatment of the cam derivative with nucleotide pyrophosphatase and alkaline phosphatase led to the liberation of 0.51 mol dephosphodi(cam)molybdopterin and of 0.59 mol guanosine per mol of enzyme, which is equal to a molar ratio of 1:2.2. The results indicate, that the organic moiety of the molybdenum cofactor of nitrate reductase from P. stutzeri is molybdopterin guanine dinucleotide of which one mol is contained per mol of nitrate reductase.  相似文献   

9.
Conditions for heterologous expression of Rhodobacter sphaeroides biotin sulfoxide reductase in Escherichia coli were modified, resulting in a significant improvement in the yield of recombinant enzyme and enabling structural studies of the molybdenum center. Quantitation of the guanine and the molybdenum as compared to that found in R. sphaeroides DMSO reductase demonstrated the presence of the bis(MGD)molybdenum cofactor. UV-visible absorption spectra were obtained for the oxidized, NADPH-reduced, and dithionite-reduced enzyme. EPR spectra were obtained for the Mo(V) state of the enzyme. X-ray absorption spectroscopy at the molybdenum K-edge has been used to probe the molybdenum coordination of the enzyme. The molybdenum site of the oxidized protein possesses a Mo(VI) mono-oxo site (Mo=O at 1.70 A) with additional coordination by approximately four thiolate ligands at 2.41 A and probably one oxygen or nitrogen at 1.95 A. The NADPH- and dithionite-reduced Mo(IV) forms of the enzyme are des-oxo molybdenum sites with approximately four thiolates at 2.33 A and two different Mo-O/N ligands at 2.19 and 1.94 A.  相似文献   

10.
Substitution of one amino acid for another at the active site of an enzyme usually diminishes or eliminates the activity of the enzyme. In some cases, however, the specificity of the enzyme is changed. In this study, we report that the changing of a metal ligand at the active site of the NiFeS-containing carbon monoxide dehydrogenase (CODH) converts the enzyme to a hydrogenase or a hydroxylamine reductase. CODH with alanine substituted for Cys(531) exhibits substantial uptake hydrogenase activity, and this activity is enhanced by treatment with CO. CODH with valine substituted for His(265) exhibits hydroxylamine reductase activity. Both Cys(531) and His(265) are ligands to the active-site cluster of CODH. Further, CODH with Fe substituted for Ni at the active site acquires hydroxylamine reductase activity.  相似文献   

11.
Self WT  Wolfe MD  Stadtman TC 《Biochemistry》2003,42(38):11382-11390
Purine hydroxylase (PH) from Clostridium purinolyticum contains a labile selenium cofactor and belongs to a class of enzymes known as the selenium-dependent molybdenum hydroxylases. The presence of approximately 1.1 mol of molybdenum, 0.87 mol of selenium, and 3.3 mol of iron per mol of PH was determined by atomic absorption spectroscopy. Enzyme preparations with lower than stoichiometric amounts of selenium exhibited correspondingly lower hydroxylase activities. Bound FAD, 1 mol per mol enzyme, was confirmed by UV-vis and fluorescence spectroscopy. CMP, released by acid hydrolysis, indicated the presence of a molybdopterin cytosine dinucleotide cofactor. The fully active PH utilized NADP(+) as an electron acceptor, and kinetic analysis revealed an optimal k(cat) of 412 s(-1) using hypoxanthine as the hydroxylase substrate. Xanthine, NAD(+), and NADPH had no significant effect on this reaction rate. A selenium-independent NADPH oxidase activity was exhibited by native PH. Electron paramagnetic resonance spectroscopy revealed the presence of a Mo(V) desulfo signal, FAD radical, and 2Fe-2S centers in hypoxanthine-reduced PH. No hyperfine coupling of selenium, using (77)Se isotope-enriched PH, was observed in any of the EPR active signals studied. The appearance of the desulfo signal suggests that the ligands of Mo in selenium-dependent molybdenum hydroxylases are different from the well-studied mammalian xanthine oxidoreductases (XOR) and aldehyde oxidoreductases (AOR) and suggests a unique role for Se in catalysis.  相似文献   

12.
The mob genes of several bacteria have been implicated in the conversion of molybdopterin to molybdopterin guanine dinucleotide. The mob locus of Rhodobacter sphaeroides WS8 comprises three genes, mobABC. Chromosomal in-frame deletions in each of the mob genes have been constructed. The mobA mutant strain has inactive DMSO reductase and periplasmic nitrate reductase activities (both molybdopterin guanine dinucleotide-requiring enzymes), but the activity of xanthine dehydrogenase, a molybdopterin enzyme, is unaffected. The inability of a mobA mutant to synthesise molybdopterin guanine dinucleotide is confirmed by analysis of cell extracts of the mobA strain for molybdenum cofactor forms following iodine oxidation. Mutations in mobB and mobC are not impaired for molybdoenzyme activities and accumulate wild-type levels of molybdopterin and molybdopterin guanine dinucleotide, indicating they are not compromised in molybdenum cofactor synthesis. In the mobA mutant strain, the inactive DMSO reductase is found in the periplasm, suggesting that molybdenum cofactor insertion is not necessarily a pre-requisite for export.  相似文献   

13.
Nelson KJ  Rajagopalan KV 《Biochemistry》2004,43(35):11226-11237
Rhodobacter sphaeroides biotin sulfoxide reductase (BSOR) contains the bis(molybdopterin guanine dinucleotide)molybdenum cofactor and catalyzes the reduction of D-biotin-D-sulfoxide to biotin. This protein is the only member of the dimethyl sulfoxide reductase family of molybdopterin enzymes that utilizes NADPH as the direct electron donor to the catalytic Mo center. Kinetic studies using stopped-flow spectrophotometry indicate that BSOR reduction by NADPH (>1000 s(-1)) is faster than steady-state turnover (440 s(-1)) and has shown that BSOR reduction occurs in concert with NADPH oxidation with no indication of a Mo(V) intermediate species. Because no crystallographic structure is currently available for BSOR, a protein structure was modeled using the structures for R. sphaeroides dimethyl sulfoxide reductase, Rhodobacter capsulatus dimethyl sulfoxide reductase, and Shewanella massilia trimethylamine N-oxide reductase as the templates. A potential NADPH-binding site was identified and tested by site-directed mutagenesis of residues within the area. Mutation of Arg137 or Asp136 reduced the ability of NADPH to serve as the electron donor to BSOR, indicating that the NADPH-binding site in BSOR is located in the active-site funnel of the putative structure where it can directly reduce the Mo center. Along with kinetic and spectroscopic data, the location of this binding site supports a direct hydride transfer mechanism for NADPH reduction of BSOR.  相似文献   

14.
l-Amino acid oxidases (LAAOs) are flavoenzymes that catalytically deaminate l-amino acids to corresponding α-keto acids with the concomitant production of ammonia (NH3) and hydrogen peroxide (H2O2). Particularly, snake venom LAAOs have been attracted much attention due to their diverse clinical and biological effects, interfering on human coagulation factors and being cytotoxic against some pathogenic bacteria and Leishmania ssp. In this work, a new LAAO from Bothrops jararacussu venom (BjsuLAAO) was purified, functionally characterized and its structure determined by X-ray crystallography at 3.1 Å resolution. BjsuLAAO showed high catalytic specificity for aromatic and aliphatic large side-chain amino acids. Comparative structural analysis with prokaryotic LAAOs, which exhibit low specificity, indicates the importance of the active-site volume in modulating enzyme selectivity. Surprisingly, the flavin adenine dinucleotide (FAD) cofactor was found in a different orientation canonically described for both prokaryotic and eukaryotic LAAOs. In this new conformational state, the adenosyl group is flipped towards the 62–71 loop, being stabilized by several hydrogen-bond interactions, which is equally stable to the classical binding mode.  相似文献   

15.
Xanthine oxidase employs four electron transport sites (flavin adenine dinucleotide (FAD), molybdenum, and two FeS centers) in catalyzing a variety of redox reactions. To determine whether the redox sites reside in independent domains of the enzyme, the temperature of heat inactivation of each site's catalytic activity was determined, except that no attempt was made to distinguish between the two FeS sites. In the oxidase form of xanthine oxidase, the order of thermal stabilities was Mo greater than FAD greater than FeS, while after conversion to its dehydrogenase form the above ranking was Mo greater than FeS greater than FAD. The small but reproducible difference in heat inactivation temperatures among the redox sites demonstrated that the sites are located in separate domains of the enzyme. To confirm the above segregation of redox centers, the temperature of heat-induced release of each redox cofactor from its site on the enzyme was examined. These temperatures were found to be different for each redox cofactor and agreed closely with the heat inactivation temperatures measured above. The data thus demonstrate that both heat inactivation and cofactor release derive from thermal unfolding of independent domains. Using a technique termed "thermal digestion analysis," the FAD domain was located in a approximately equal to 42,000-Da tryptic fragment, while the FeS and Mo domains were isolated in a trypsin-resistant 92,000-Da fragment. We conclude that xanthine oxidase is constructed in modular fashion with the redox sites located in independent structural domains.  相似文献   

16.
17.
《BBA》2014,1837(2):277-286
In Rhodobacter sphaeroides periplasmic nitrate reductase NapAB, the major Mo(V) form (the “high g” species) in air-purified samples is inactive and requires reduction to irreversibly convert into a catalytically competent form (Fourmond et al., J. Phys. Chem., 2008). In the present work, we study the kinetics of the activation process by combining EPR spectroscopy and direct electrochemistry. Upon reduction, the Mo (V) “high g” resting EPR signal slowly decays while the other redox centers of the protein are rapidly reduced, which we interpret as a slow and gated (or coupled) intramolecular electron transfer between the [4Fe–4S] center and the Mo cofactor in the inactive enzyme. Besides, we detect spin–spin interactions between the Mo(V) ion and the [4Fe–4S]1 + cluster which are modified upon activation of the enzyme, while the EPR signatures associated to the Mo cofactor remain almost unchanged. This shows that the activation process, which modifies the exchange coupling pathway between the Mo and the [4Fe–4S]1 + centers, occurs further away than in the first coordination sphere of the Mo ion. Relying on structural data and studies on Mo-pyranopterin and models, we propose a molecular mechanism of activation which involves the pyranopterin moiety of the molybdenum cofactor that is proximal to the [4Fe–4S] cluster. The mechanism implies both the cyclization of the pyran ring and the reduction of the oxidized pterin to give the competent tricyclic tetrahydropyranopterin form.  相似文献   

18.
Mononuclear Mo-containing enzymes of the xanthine oxidase (XO) family catalyze the oxidative hydroxylation of aldehydes and heterocyclic compounds. The molybdenum active site shows a distorted square-pyramidal geometry in which two ligands, a hydroxyl/water molecule (the catalytic labile site) and a sulfido ligand, have been shown to be essential for catalysis. The XO family member aldehyde oxidoreductase from Desulfovibrio gigas (DgAOR) is an exception as presents in its catalytically competent form an equatorial oxo ligand instead of the sulfido ligand. Despite this structural difference, inactive samples of DgAOR can be activated upon incubation with dithionite plus sulfide, a procedure similar to that used for activation of desulfo-XO. The fact that DgAOR does not need a sulfido ligand for catalysis indicates that the process leading to the activation of inactive DgAOR samples is different to that of desulfo-XO. We now report a combined kinetic and X-ray crystallographic study to unveil the enzyme modification responsible for the inactivation and the chemistry that occurs at the Mo site when DgAOR is activated. In contrast to XO, which is activated by resulfuration of the Mo site, DgAOR activation/inactivation is governed by the oxidation state of the dithiolene moiety of the pyranopterin cofactor, which demonstrates the non-innocent behavior of the pyranopterin in enzyme activity. We also showed that DgAOR incubation with dithionite plus sulfide in the presence of dioxygen produces hydrogen peroxide not associated with the enzyme activation. The peroxide molecule coordinates to molybdenum in a η2 fashion inhibiting the enzyme activity.  相似文献   

19.
The paramagnetic iron at the active site of highly purified, catalytically active phenylalanine hydroxylase was studied by EPR at 3.6 K and one-dimensional 1H-NMR spectroscopy at 293 K. The EPR-detectable iron of the bovine enzyme was found to be present as a high-spin form (S = 5/2) in different ligand field symmetries depending on medium conditions (buffer ions) and the presence of ligands known to bind at the active site. At 3.6 K and in phosphate buffer, the paramagnetic iron is coordinated in an environment of rhombic symmetry (g = 4.3), whereas Tris buffer favours an environment of axial ligand field symmetry (g = 6.7, 5.3 and 2.0). The latter axial type of signals resembles those observed at g = 7.0, 5.2 and 1.9 for the enzyme in phosphate buffer when L-noradrenaline is added as an active-site ligand (inhibitor). The same proportion of iron that coordinates to L-noradrenaline seems to be reduced by the pterin cofactor and participate in catalysis. Experimental evidence is presented that Tris inhibits the enzyme by interacting with the enzyme-bound ferric iron and decreases its rate of reduction by the tetrahydropterin cofactor. Preincubation with dithiothreitol also inhibits the enzyme activity and prevents the reduction of its catalytically active ferric iron by pterin cofactors as well as binding of catecholamines to the enzyme. 1H-NMR spectroscopy revealed that the substrate (L-phenylalanine) and L-noradrenaline bind close to the paramagnetic iron, and that the catecholamine displaces the substrate from its binding at the active site. The results support our recently proposed model for the cooperative binding of inhibitor and substrate at the active site [Martínez, A. et al. (1990) Eur. J. Biochem. 193, 211-219].  相似文献   

20.
The periplasmic nitrate reductase from Paracoccus denitrificans is a soluble two-subunit enzyme which binds two hemes (c-type), a [4Fe-4S] center, and a bis molybdopterin guanine dinucleotide cofactor (bis-MGD). A catalytic cycle for this enzyme is presented based on a study of these redox centers using electron paramagnetic resonance (EPR) and extended X-ray absorption fine structure (EXAFS) spectroscopies. The Mo(V) EPR signal of resting NAP (High g [resting]) has g(av) = 1.9898 is rhombic, exhibits low anisotropy, and is split by two weakly interacting protons which are not solvent-exchangeable. Addition of exogenous ligands to this resting state (e.g., nitrate, nitrite, azide) did not change the form of the signal. A distinct form of the High g Mo(V) signal, which has slightly lower anisotropy and higher rhombicity, was trapped during turnover of nitrate and may represent a catalytically relevant Mo(V) intermediate (High g [nitrate]). Mo K-edge EXAFS analysis was undertaken on the ferricyanide oxidized enzyme, a reduced sample frozen within 10 min of dithionite addition, and a nitrate-reoxidized form of the enzyme. The oxidized enzyme was fitted best as a di-oxo Mo(VI) species with 5 sulfur ligands (4 at 2. 43 A and 1 at 2.82 A), and the reduced form was fitted best as a mono-oxo Mo(IV) species with 3 sulfur ligands at 2.35 A. The addition of nitrate to the reduced enzyme resulted in reoxidation to a di-oxo Mo(VI) species similar to the resting enzyme. Prolonged incubation of NAP with dithionite in the absence of nitrate (i.e., nonturnover conditions) resulted in the formation of a species with a Mo(V) EPR signal that is quite distinct from the High g family and which has a g(av) = 1.973 (Low g [unsplit]). This signal resembles those of the mono-MGD xanthine oxidase family and is proposed to arise from an inactive form of the nitrate reductase in which the Mo(V) form is only coordinated by the dithiolene of one MGD. In samples of NAP that had been reduced with dithionite, treated with azide or cyanide, and then reoxidized with ferricyanide, two Mo(V) signals were detected with g(av) elevated compared to the High g signals. Kinetic analysis demonstrated that azide and cyanide displayed competitive and noncompetitive inhibition, respectively. EXAFS analysis of azide-treated samples show improvement to the fit when two nitrogens are included in the molybdenum coordination sphere at 2.52 A, suggesting that azide binds directly to Mo(IV). Based on these spectroscopic and kinetic data, models for Mo coordination during turnover have been proposed.  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