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 62 毫秒
1.
Interaction of dioxouranium(VI) (uranyl) ion with ATP was studied by ligand/proton and metal/hydroxide displacement technique, at very low ionic strength and at I=0.15 mol L(-1), in aqueous Me4NCl and NaCl solutions, at t=25 degrees C. Measurements were carried out in the pH range 3-8.5, before the formation of precipitate. Computer analysis allowed us to find the quite stable species UO2(ATP)H2(0), UO2(ATP)H-, UO2(ATP)2-, UO2(ATP)2(6-), UO2(ATP)2H2(4-) and UO2(ATP)(OH)3- whose formation constants are (at I=0 mol L(-1)) logbeta(112)=18.21, logbeta(111)=14.70, logbeta(110)=9.14, logbeta(120)=12.84, logbeta(122)=24.82, and logbeta(11-1)=2.09, respectively. Different values were obtained in the above ionic media at I=0.15 mol L(-1) and the dependence on the ionic medium was interpreted in terms of interactions between the negatively charged complex species and cations of supporting electrolytes. The species more stable in NaCl than in Me4NCl are those with the highest negative charge, UO2(ATP)2(6-) and UO2(ATP)2H2(4-), and the extra stability of these species can be attributed to the interaction with Na+. Speciation profiles show that ATP can suppress UO2(2+) hydrolysis, and that in the neutral to slightly alkaline range the yield of complex UO2-ATP species is quite high. Comparison with other metal-ATP systems is also given in order to recognize the possibility of binding competition of uranyl ion in metal-ATP requiring enzymes for biochemical processes.  相似文献   

2.
Complexes of the uranyl cation [UO(2)(2+)] with histidine (His), N-acetyl-histidine (NAH), tyrosine (Tyr), and N-acetyl-tyrosine (NAT) were studied by UV-visible and NMR spectroscopy, and by potentiometric titration. Protonation constants for each ligand are reported, as are cumulative formation constants for uranyl-amino acid complexes. Coupling constant data (J(CH)) for uranyl-histidine complexes indicate that inner-sphere solution interactions between histidine and uranyl cation are solely at the carboxylate site. At 25 degrees C the major uranyl-histidine complex has a cumulative formation constant of logbeta(110)=8.53, and a proposed formula of [UO(2)HisH(2)(OH)(2)](+); the stepwise formation constant, logK(UL), is estimated to be 5.6 ( approximately 8.53-(-6.1)-(-6.1)-15.15). Outer-sphere interactions, H-bonding or electrostatic interactions, are proposed as contributing a significant portion of the stability to the ternary uranyl-hydroxo-amino acid complexes. The temperature dependent protonation constants of histidine and formation constants between uranyl cation and histidine are reported from 10 to 35 degrees C; at 25 degrees C, DeltaG=-43.3 kJ/mol.  相似文献   

3.
A derivative of 1,10-phenanthroline that binds to UO(2)(2+) with nanomolar affinity was found to be a very effective immunogen for the generation of antibodies directed toward chelated complexes of hexavalent uranium. This study describes the synthesis of 5-isothiocyanato-1,10-phenanthroline-2,9-dicarboxylic acid and its use in the generation and functional characterization of a group of monoclonal antibodies that recognize the most soluble and toxic form of uranium, the hexavalent uranyl ion (UO(2)(2+)). Three different monoclonal antibodies (8A11, 10A3, and 12F6) that recognize the 1:1 complex between UO(2)(2+) and 2,9-dicarboxy-1,10-phenanthroline (DCP) were produced by the injection of BALB/c mice with DCP-UO(2)(2+) covalently coupled to a carrier protein. Equilibrium dissociation constants for the binding of DCP-UO(2)(2+) to antibodies 8A11, 10A3, and 12F6 were 5.5, 2.4, and 0.9 nM, respectively. All three antibodies bound the metal-free DCP with roughly 1000-fold lower affinity. The second-order rate constants for the bimolecular association of each antibody with soluble DCP-UO(2)(2+) were in the range of 1 to 2 x 10(7) M(-1) s(-1). Binding studies conducted with structurally related chelators and 21 metal ions demonstrated that each of these three antibodies was highly specific for the soluble DCP-UO(2)(2+) complex. Detailed equilibrium binding studies conducted with three other derivatives of DCP, either complexed with UO(2)(2+) or metal-free, suggested that the antigen binding sites on the three antibodies have significant functional and structural similarities. Biomolecules that bind specifically to uranium will be at the heart of any new biotechnology developed to monitor and control uranium contamination. The three antibodies described herein possess sufficient affinity and specificity to support the development of immunoassays for hexavalent uranium in environmental and clinical samples.  相似文献   

4.
Ion exchange/complexation of the uranyl ion by Rhizopus biosorbent   总被引:3,自引:0,他引:3  
Nonliving biomass of nine Rhizopus species effectively sequestered the uranyl ion from solution, taking up 150-250 mg U/g dry cells at 300 ppm U equilibrium concentration in solution, and 100-160 mg U/g dry cells with 100 ppm U in solution. The affinity of this biosorbent for the uranyl ion was found to be affected by timing of harvesting and medium composition. Uptake of the uranyl ion by nonliving biomass of Rhizopus oligosporus was due to ion exchange or complexation, since binding was reversed by the addition of complexing ligands or the reduction of pH to a value less than 2. Uptake isotherms were interpreted in terms of a model of multiple equilibria. At pH 相似文献   

5.
The polyphosphate kinase gene from Pseudomonas aeruginosa was overexpressed in its native host, resulting in the accumulation of 100 times the polyphosphate seen with control strains. Degradation of this polyphosphate was induced by carbon starvation conditions, resulting in phosphate release into the medium. The mechanism of polyphosphate degradation is not clearly understood, but it appears to be associated with glycogen degradation. Upon suspension of the cells in 1 mM uranyl nitrate, nearly all polyphosphate that had accumulated was degraded within 48 h, resulting in the removal of nearly 80% of the uranyl ion and >95% of lesser-concentrated solutions. Electron microscopy, energy-dispersive X-ray spectroscopy, and time-resolved laser-induced fluorescence spectroscopy (TRLFS) suggest that this removal was due to the precipitation of uranyl phosphate at the cell membrane. TRLFS also indicated that uranyl was initially sorbed to the cell as uranyl hydroxide and was then precipitated as uranyl phosphate as phosphate was released from the cell. Lethal doses of radiation did not halt phosphate secretion from polyphosphate-filled cells under carbon starvation conditions.  相似文献   

6.
The effect of different extracellular alkaline-earth cations (Ca2+, Mg2+, Sr2+, Ba2+) upon the threshold membrane potential for spike initiation in crayfish axon has been studied by means of intracellular microelectrodes. This was done at the following extracellular concentrations of the divalent uranyl ion (UO2/2+): 1.0 X 10(-6) M, 3.0 X 10(-6) M, and 9.0 X 10(-6) M. At each concentration employed, extensive neutralization of axonal surface charges by UO2/2+ was evidenced by the fact that equal concentrations (50 mM) of alkaline-earth cations did not have the same effect on the threshold potential. The selectivity sequences observed at the different uranyl-ion concentrations were: 1.0 X 10(-6) M UO2/2+, Ca2+ greater than Mg2+ greater than Sr2+ greater than Ba2+; 3.0 X 10(-6) M UO2/2+, Ca2+ greater than Mg2+ greater than Ba2+ larger than or equal to Sr2+; 9.0 X 10(-6) M UO2/2+, Ca2+ approximately Ba2+ greater than Sr2+ greater than Mg2+. These selectivity sequences are in accord with the equilibrium selectivity theory for alkaline-earth cations. At each of the concentrations used, uranyl ion did not have any detectable effect on the actual shape of the action potential itself. It is concluded that many (if not most) of the surface acidic groups in the region of the sodium gates represent phosphate groups of membrane phospholipids, but that the m gates themselves are probably protein-aceous in structure.  相似文献   

7.
The polyphosphate kinase gene from Pseudomonas aeruginosa was overexpressed in its native host, resulting in the accumulation of 100 times the polyphosphate seen with control strains. Degradation of this polyphosphate was induced by carbon starvation conditions, resulting in phosphate release into the medium. The mechanism of polyphosphate degradation is not clearly understood, but it appears to be associated with glycogen degradation. Upon suspension of the cells in 1 mM uranyl nitrate, nearly all polyphosphate that had accumulated was degraded within 48 h, resulting in the removal of nearly 80% of the uranyl ion and >95% of lesser-concentrated solutions. Electron microscopy, energy-dispersive X-ray spectroscopy, and time-resolved laser-induced fluorescence spectroscopy (TRLFS) suggest that this removal was due to the precipitation of uranyl phosphate at the cell membrane. TRLFS also indicated that uranyl was initially sorbed to the cell as uranyl hydroxide and was then precipitated as uranyl phosphate as phosphate was released from the cell. Lethal doses of radiation did not halt phosphate secretion from polyphosphate-filled cells under carbon starvation conditions.  相似文献   

8.
Biosorption of actinides like uranium by fungal cells can play an important role in the mobilization or immobilization of these elements in nature. Sorption experiments of U(VI) with Schizophyllum commune at different initial uranium concentrations and varying metal speciation showed high uranium sorption capacities in the pH range of 4–7. A combination of high angle annular dark-field and scanning transmission electron microscopy analysis (HAADF-STEM) showed that living mycelium cells accumulate uranium at the cell wall and intracellular. For the first time the fluorescence properties of uranium accumulates were investigated by means of time-resolved laser-induced fluorescence spectroscopy (TRLFS) beside the determination of corresponding structural parameters using X-ray absorption fine structure spectroscopy (EXAFS). While the oxidation state of uranium remained unchanged during sorption, uranium speciation changed significantly. Extra and intracellular phosphate groups are mainly responsible for uranium binding. TRLFS spectra clearly show differences between the emission properties of dissolved species in the initial mineral medium and of uranium species on fungi. The latter were proved to be organic and inorganic uranyl phosphates formed depending on the uranyl initial concentration and in some cases on pH.  相似文献   

9.
The vitamin B(6) derivatives 4-pyridoxic acid (anionic) and the Schiff base N,N'-ethylenebis(pyridoxylideneiminato) react with UO(2)(NO(3))(2) * 6H(2)O to give [UO(2)(beta-pyracinide)(2)(H(2)O)] (beta-pyracin=4-pyridoxic acid) and [UO(2)(Pyr(2)en)DMSO]Cl(2)(Pyr(2)en=N,N'-ethylenebis(pyridoxylideneiminato); DMSO=dimethyl sulfoxide). In both compounds the two uranyl oxo ligands set the axis of distorted pentagonal bipyramides. The ability of vitamin B(6) derivatives to react with UO(2)(2+) allowing the chelation of one uranium atom represents a very specific model of assimilation of uranium by living beings. It could also explain the serious damages caused by heavy or radioactive metals like uranium since their complexation "in vivo" by enzymatic systems like pyridoxal phosphate-containing enzymes would lead to a modification of the prosthetic groups of the metalloenzymes with loss of their catalytic activities.  相似文献   

10.
Time-resolved laser-induced fluorescence spectroscopy (TRLFS) was used to study the properties of uranium complexes (emission spectra and fluorescence lifetimes) formed by the cells of the three recently described eco-types of Acidithiobacillus ferrooxidans. The results demonstrated that these complexes have different lifetimes which increase in the same order as the capability of the strains to accumulate uranium. The complexes built by the cells of the eco-type II were the strongest, whereas, those of the eco-types I and III were significantly weaker. The emission spectra of all A. ferrooxidans complexes were almost identical to those of the uranyl organic phosphate compounds. The latter finding was confirmed by infrared spectroscopic analysis.  相似文献   

11.
Four imidazolate-bridged binuclear copper(II)-copper(II) and copper(II)-zinc(II) complexes viz., [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH, [(Phen)(2)Cu-Im-Cu(Phen)(2)](BF(4))(3).2CH(3)OH, [(Bipy)(2)Cu-Im-Zn(Bipy)(2)](BF(4))(3), and [(Phen)(2)Cu-Im-Zn(Phen)(2)](BF(4))(3), (Bipy=2,2'-Bipyridyl, Phen=1-10-Phenanthroline and Im=imidazolate ion) were synthesized as a possible models for superoxide dismutase (SOD). Complex [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH has been structurally characterized. This complex crystallizes in the triclinic space group P1, with the unit parameters a=8.88(5) A, b=13.79(17) A, c=20.18(18) A, alpha=76.424(8)(o), beta=85.888(6)(o), gamma=82.213(7). The metal-nitrogen bond length from 1.972-2.273 A and the distance Cu-Cu is 5.92 A. The five-coordinate geometry about the copper(II) ion is square pyramidal. Magnetic moment and electron paramagnetic resonance (e.p.r.) spectral measurements of the homobinuclear complexes have shown an antiferromagnetic exchange interaction. From the e.p.r. and UV-Vis spectral measurement studies, these complexes have been found to be stable (pH 8.5-10.5 for 1, 10.5 for 2,3 and 8.5 for 4). These complexes catalyse the dismutation of superoxide radical (O(2)(-)) at biological pH. All the observations indicate that these complexes act as good possible models for superoxide dismutase.  相似文献   

12.
Abstract A study of cell-wall site interaction of uranyl ion adsorbed by non-proliferative suspensions of Mycobacterium smegmatis at pH 1 has been carried out using extracts of arabinogalactan-peptidoglycan and phospholipids obtained from whole cells treated under sorption conditions. Evidence for binding of UO22+ by constituent P-lipids was provided by comparative 31P-NMR and IR spectroscopic measurements of the isolated wall fractions and of the model complex uranyl-phosphatidyl inositol.  相似文献   

13.
The complexation of aluminium(III) with glycerol-1-phosphate (G1P) and glycerol-2-phosphate (G2P) in aqueous solutions has been studied as a function of pH, by pH-potentiometry, 31P NMR spectroscopy and ESI mass spectrometry. Various mononuclear complexes (MLH(2)(3+), MLH(2+), ML(+), ML(2)H, ML(2)(-)) and polynuclear species (M(3)L(3)H(-1)(2+), M(3)L(2)H(-n)((n-5)-) with n=5, 6, 7, M(2)L(2)H(-1)(+) ) are formed in the system where the full protonated ligands are noted LH(2). NMR experiments clearly show that G1P and G2P already interact with Al(III) at pH 1. The potentiometric results are confirmed by ESI measurements and 31P NMR studies. No metal ion-induced deprotonation and coordination of the alcoholic-OH functions seem to occur during the complexation. The situation is very different for the glycerophosphorylcholine ligand (GPC identical with LH). Only the complex ML(3+) is formed in aqueous solution with a relatively low formation constant (K=5 at 37 degrees C). This species is clearly identified in 31P and 27Al NMR spectra. The complexation study as a function of the temperature allowed us to determine the thermodynamic parameters of the complex formation. The complexation is not governed by the reaction enthalpy that is found to be positive but by the entropy that is largely positive.  相似文献   

14.
1. Reactions between triphosphoinositide and the basic experimental allergic encephalitogenic (EAE) protein were examined in aqueous solution and in a biphasic solvent system (chloroform-methanol-water, 8:4:3, by vol.). 2. In the absence of salt an insoluble complex (I) is formed containing triphosphoinositide and EAE protein in proportions that represent complete neutralization of lipid and protein at the pH concerned. 3. In the presence of a low concentration (0.05m) of sodium chloride an insoluble positively charged complex (II) forms. It contains triphosphoinositide and EAE protein in a lower concentration ratio than complex I. This complex, which has a constant composition between pH7.5 and pH10, can take up additional micellar triphosphoinositide producing complex I, which can then be solubilized by excess of triphosphoinositide. 4. The complexes are dissociated by more concentrated sodium chloride solutions and low concentrations of calcium chloride, suggesting that they are largely stabilized by electrostatic bonds. The protein recovered after dissociation is immunologically active and has the same electrophoretic mobility as the original. 5. Water-insoluble ternary complexes containing triphosphoinositide, EAE protein and large amounts of phosphatidylcholine can be prepared. From these, chloroform-methanol (2:1, v/v) extracts only phosphatidylcholine. 6. An insoluble ternary complex of Ca(2+) ion, EAE protein and triphosphoinositide can be prepared by adding calcium chloride to a complex I preparation solubilized by excess of triphosphoinositide. 7. EAE protein will also form complexes with other acidic phospholipids, e.g. phosphatidic acid, phosphatidylserine and phosphatidylinositol, but not with phosphatidylcholine or phosphatidylethanolamine. The phosphatidylinositol and phosphatidylserine complexes are chloroform soluble, i.e. proteolipids. 8. The possibility that complexes between EAE protein and acidic phospholipids occur in vivo is discussed. Triphosphoinositide and EAE protein occur in ox brain myelin in approximately the same concentration ratios as they do in complex II, formed at physiological salt concentration and pH.  相似文献   

15.
Hydrogen uranyl phosphate (HUO2PO4: HUP), deposited enzymatically on Citrobacter N14 cells immobilized as biofilm on ceramic Raschig rings in a flow-through column, removed nickel quantitatively from dilute aqueous solution in the form of nickel uranyl phosphate, via intercalative ion exchange. Nickel-loaded columns were regenerated by washing either with citrate buffer or with buffer containing UO22 and phosphate donor (glycerol 2-phosphate), this giving additional crystalline HUP deposit for subsequent improvement of nickel removal. No uranium release occurred during selective desorption of Ni, proving the integrity of the biofilm within the column. The use of ceramic supports to manufacture an artificial, bioinorganic, ion exchanger is novel and the use of solid matrices overcomes the problems of mechanical stability which limit the applications of gel-immobilized cells for large-scale processes. © Rapid Science Ltd. 1998  相似文献   

16.
The interaction of UO2(2+) with dipalmitoylphosphatidylcholine (DPPC) has been studied as a function of temperature and composition using nuclear magnetic resonance (NMR) spectroscopy, differential scanning calorimetry (DSC), and monolayer studies. Computer simulations of the 31P-NMR powder spectra of DPPC dispersions in the presence of various concentrations of UO2(2+) are consistent with the binding stoichiometry of [UO2(2+)]/[DPPC] = 1:4 at [UO2(2+)]/[DPPC] less than 0.3. This complex undergoes a phase transition to the liquid crystalline phase at T'm = 50 +/- 3 degrees C with a breadth delta T'm = 7 +/- 3 degrees C. This broad transition gradually disappears at higher UO2(2+) concentrations, suggesting the presence of yet another UO2(2+)/DPPC complex (or complexes) whose NMR spectra are indistinguishable from those of the 1:4 UO2(2+)/DPPC species. The temperature-dependent 13C powder spectra of 2(1-13C) DPPC dispersions in the presence of 1.2 mol ratio of UO2(2+) show that this higher order complex (complexes) also undergoes a phase transition to the liquid crystalline state at T'm +/- = 58 +/- 3 degrees C with a breadth delta T"m = 15 +/- 5 degrees C. The NMR spectra indicate that exchange among these various UO2(2+)/DPPC complexes is slow. In addition, computer simulations of the 31P-, 13C-, and 2H-NMR powder spectra show that axial diffusion of the DPPC molecules about their long axes is quenched by addition of UO2(2+) and acyl chain isomerization is the dominant motional mode. The isomerization is best described as two-site hopping of the greater than C-D bond at a rate of approximately 10(6) s-1, a motional mode which is expected for a kink diffusion.  相似文献   

17.
The energy landscape of the uranyl (UO2) chelate dissociated from a monoclonal antibody U08S was investigated using dynamic force spectroscopy (DFS). The uranyl ion (UO2(2+)) is chelated with the ligand dicarboxy-phenanthroline (DCP). The monoclonal antibody U08S was raised against UO2-DCP and does not cross-react with DCP alone. The results of plotting the most probable force against the logarithm of the loading rate show two distinguished values of slopes of multiple fitting lines, as observed in our previous study on that system with monoclonal antibody U04S (Odorico et al., 2007a. Biophys. J. 93: 645-654.). It indicates an unbinding process undergoing at least two activation states. We have generated the histogram of unbinding events with respect to the composite stiffness of the complex between the protein and the uranyl compound. Combining the model of Bell and Evans with that of Williams, we have estimated the number of parallel bonds involved in the unbinding process and determined the value of stiffness for individual bonds. We propose that the uranyl compound binds to the two antibodies U04S and U0c at structurally equivalent locations and forms the interaction with similar coordination modes. In addition, the unbinding process goes through two steps; the first weakens the bonding of the central metal with AspL50 of the antibody and the second breaks other non-bonded interactions of the compound with the antibody.  相似文献   

18.
Fungal transformations of uranium oxides   总被引:1,自引:0,他引:1  
The biogeochemical activities of free-living and symbiotic fungi must be acknowledged in attempts to understand uranium cycling and dispersal in the environment. Although the near-surface geochemistry of uranium is very complex and a wide variety of mineral phases is known, uranium trioxide (UO3) and triuranium octaoxide (U(3)O(8)) can be used as well characterized models in the study of biotransformations. We have used a complex methodological approach involving advanced solid state speciation and scanning electron microscopy to study the ability of saprotrophic, ericoid and ectomycorrhizal fungi to transform these model oxides. This study has revealed that fungi exhibit a high uranium oxide tolerance, and possess the ability to solubilize UO3 and U(3)O(8) and to accumulate uranium within the mycelium to over 80 mg (g dry weight)(-1) biomass. X-ray absorption spectroscopy of uranium speciation within the biomass showed that in most of the fungi the uranyl ion was coordinated to phosphate ligands, but in ectomycorrhizal fungi mixed phosphate/carboxylate coordination was observed. Abundant uranium precipitates associated with phosphorus were found in the mycelium and encrusted the hyphae. Some of the fungi caused the biomineralization of well-crystallized uranyl phosphate minerals of the meta-autunite group. This is the first experimental evidence for fungal transformations of uranium solids and the production of secondary mycogenic uranium minerals.  相似文献   

19.
The binuclear cobalt complex [Co(2)(Me(2)dtc)(5)](+) reacts with a range of nitrogen donor ligands L' or L' to form an equimolar mixture of Co(Me(2)dtc)(3) and the mixed-ligand complexes [Co(Me(2)dtc)(2)(L')(2)](+) or [Co(Me(2)dtc)(2)(L')](+), where (L')(2) is two monodentate ligands and (L') is one bidentate ligand. The complexes prepared by this route contain the monodentate ligands L'=1-methyl-imidazole, 1-methyl-5-nitro-imidazole and benzimidazole, all of which coordinate to cobalt through an imidazole nitrogen atom. Symmetrical bidentate ligand complexes contain the bisimidazole L'=2,2'-bis(4,5-dimethylimidazole), the diamine L'=1,2-diaminobenzene and the pyridine donors L'=2,2'-bipyridine, 4,4'-dimethyl-2,2'-bipyridine and 1,10-phenanthroline. Two examples of complexes with unsymmetrical bidentate imidazole-amine donors were prepared in which L'=4-(2-aminoethyl)imidazole (histamine) and 2-aminomethylbenzimidazole. All new complexes were fully characterised, and the X-ray crystal structure of the histamine complex [Co(Me(2)dtc)(2)(hist)]ClO(4) is also reported.  相似文献   

20.
Yuan F  Chen X  Liu Y  Zhang T  Sun D  Liu J 《Chirality》2012,24(2):174-180
In this study, two isomeric ruthenium(II) complexes [Ru(bpy)(2)(p-mopip)](2+) (1) and [Ru(bpy)(2)(o-mopip)](2+) (2) (bpy = 2, 2-bipyridine; L: p-mopip = 2-(4-methoxylphenyl) imidazo [4,5-f][1,10]phenanthroline, o-mopip = 2-(2-methoxylphenyl) imidazo[4,5-f][1,10] phenan-throline) contained -OCH(3) at different positions on the phenyl ring and their enantiomers Λ-1, -2 and Δ-1, -2 displayed different properties. The cell viability of these ruthenium(II) complexes was evaluated by MTT, and complex Λ-1 has shown significant higher anticancer potency than Δ-1 against all the cell lines screened. Fluorescence microscopy and flow cytometric analyses demonstrated that complex Λ-1 was able to induce apoptosis. The interactions of complexes Λ-1, 1, and Δ-1 with bovine serum albumin (BSA) were investigated by fluorescence and circular dichroism (CD) measurements. The fluorescence quenching mechanism of BSA by complexes Λ-1, 1, and Δ-1 was determined to be a static process, and the apparent binding constant K(a) values is as follows: Λ-1 >1 > Δ-1. The number of binding sites n for all these complexes was 1. The result of CD showed that the secondary structure of BSA molecules was changed in the presence of the ruthenium(II) complex.  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