首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A 191Os→ 191mIr generator has been developed that has higher 191mIr yield and lower 191Os breakthrough than previous designs. These improvements have been realized through the use of the osmium chelate complex trans-dioxobismalonatoosmate(VI) as the parent species on the generator. The new generator provides an initial 191mIr yield of 40%/mL and 191Os breakthrough of 2–3 × 10−3% when eluted with a solution of 0.05 M malonic acid/0.10 M sodium chloride at pH 4. Other advantages of the new design include faster clearance of the 191Os breakthrough products and simpler assembly.  相似文献   

2.
《Inorganica chimica acta》1988,146(2):187-191
Lithium penta(cyano-13C)nitrosylruthenate (2-), Li2[Ru(13CN)5NO], in which the anion is the ruthenium analogue of the nitroprusside ion, has been synthesized at 90% isotopic enrichment, and characterized spectroscopically. Despite the very high level of 13C enrichment, no two-bond coupling 2J(13Cax-Ru13Ceq) was detected in the high-frequency 13C NMR spectrum of Li2[Ru(13CN)5NO], nor was any such coupling observed in Li4[Ru(13CN)5(15NO2)] although both two-bond couplings to 15N, 2J(13Cax-Ru15NO2) and 2J(13CeqRu15N) were observed. Li2[Ru(13CN)5(14NO)] reacted with excess of Li[15NO2] to yield Li4[Ru(13CN)5(15NO2)] only: no Li2[Ru(13CN)5(15NO)] was observed. Li4[Ru(13CN)5(14NO2)] however showed no exchange with Li[15NO2]. While [Ru(CN)5NO]2− reacted with both OH and SH in reactions similar to those of [Fe(CN)5NO]2−, no reactions were detected between [Ru(CN)5NO]2− and piperidine, [CH(CN)2], [CH(COCH3)2], MeS, or [S2O4]2−, all of which are known to react readily with [Fe(CN)5NO]2−  相似文献   

3.
Abstract. Nitrate uptake into Chara corallina cells at different external pH (pHo) after different NO3 pretreatment conditions has been investigated. Following NO3 pretreatment (0.2 mol m−3 NO3) there was little effect of pHo on subsequent net NO3 uptake into Chara cells. After N deprivation (2 mmol m−3 NO3) there was a pronounced effect of pHo on nitrate uptake, the maximum rate occurring at pHo 4.7. There was no consistent relationship between OH efflux and NO3 uptake in short term experiments (< 1 h). NO3 efflux was also sensitive to pHo, the maximum rate occurring at ∼ pHo 5.0. An inhibitor of the proton pump, DES, immediately stimulated NO3 uptake into cells pretreated with NO3 and prevented the time-dependent decrease in NO3, uptake into Chara cells that had been previously treated with low N (2 mmol m−3 NO3). NO3 efflux was found to be very sensitive to DES with Ki= 0.7 mmol m−3. At the optimum pHo for NO3 uptake the effect of DES on membrane potential (ψm) were slight, and only apparent after 30 min. The results are interpreted in context of current models relating NO3 uptake and H+ pump activity. A new model for NO3 uptake is proposed which involves NO3/NO3 exchange at steady state.  相似文献   

4.
《Inorganica chimica acta》1987,136(3):177-183
Formation constants for cadmium(II) complexes of mercaptosuccinate in aqueous solution have been determined at 37 °C and 150 mmol dm−3 chloride medium by potentiometric titrations using a glass electrode. The emf data obtained have been analyzed using the ESTA computer program library. The experimental data can be explained by the formation of the complexes Cd3(MSA)2, Cd2(MSA)2H, Cd2- (MSA)22−, Cd2(MSA)2OH3− and Cd(MSA)24−. The formation constants obtained have been used in a simulation model of blood plasma to investigate the mobilization of cadmium(II) by mercaptosuccinate in normal plasma. It is shown that mobilization is unlikely at pharmacological levels of the drug.  相似文献   

5.
Binary and ternary systems involving adenosine 5′-triphosphate (ATP), 2,2′-dipyridylamine (DPA) and magnesium, calcium, strontium, manganese, cobalt, copper, and zinc(II) metal ions have been investigated in aqueous media by potentiometric titrations. The analysis of the titration curves shows the existence of M(ATP)2−, M(ATP)(H), and M(ATP)2(H)24− species for alkaline-earth metal ions, while no ternary complex can be detected. For transition metal ions both binary and ternary species are found. Binary M(ATP)2(H)24− complexes are present in solutions containing manganese and cobalt(II) metal ions but these species cannot be revealed in the case of copper and zinc(II). Ternary complexes as M(ATP)(DPA)2− and M(ATP)(DPA)(H) are common to all transition metals. Binuclear and hydroxo complexes as M2(ATP)(OH) and M(ATP)(OH)3− are found only for copper and zinc(II). A hypothesis on the possible role of the species M-ATP in 1:2 ratio in the dephosphorylation mechanism is advanced on the basis of a comparison between the equilibrium data in the solution phase and the solid state structures of the magnesium, calcium, and manganese(II)- ATP-DPA systems.  相似文献   

6.
Nitric oxide (NO) is a chemical weapon within the arsenal of immune cells, but is also generated endogenously by different bacteria. Pseudomonas aeruginosa are pathogens that contain an NO-generating nitrite (NO2) reductase (NirS), and NO has been shown to influence their virulence. Interestingly, P. aeruginosa also contain NO dioxygenase (Fhp) and nitrate (NO3) reductases, which together with NirS provide the potential for NO to be metabolically cycled (NO→NO3→NO2→NO). Deeper understanding of NO metabolism in P. aeruginosa will increase knowledge of its pathogenesis, and computational models have proven to be useful tools for the quantitative dissection of NO biochemical networks. Here we developed such a model for P. aeruginosa and confirmed its predictive accuracy with measurements of NO, O2, NO2, and NO3 in mutant cultures devoid of Fhp or NorCB (NO reductase) activity. Using the model, we assessed whether NO was metabolically cycled in aerobic P. aeruginosa cultures. Calculated fluxes indicated a bottleneck at NO3, which was relieved upon O2 depletion. As cell growth depleted dissolved O2 levels, NO3 was converted to NO2 at near-stoichiometric levels, whereas NO2 consumption did not coincide with NO or NO3 accumulation. Assimilatory NO2 reductase (NirBD) or NorCB activity could have prevented NO cycling, and experiments with ΔnirB, ΔnirS, and ΔnorC showed that NorCB was responsible for loss of flux from the cycle. Collectively, this work provides a computational tool to analyze NO metabolism in P. aeruginosa, and establishes that P. aeruginosa use NorCB to prevent metabolic cycling of NO.  相似文献   

7.
A novel ruthenium(II) complex of dipyridophenazine (DPPZ) with the ancillary ligand imidazole[4,5-f] [1,10]phenanthroline (IP), [Ru(IP)2(DPPZ)] (PF6)2, has been synthesized and characterized by elemental analysis, 1D and 2D 1H NMR, fast-atom bombardment mass spectra (FABMS), electronic spectroscopy and cyclic voltammetry. The DNA-binding properties of the complex were studied by spectroscopic methods. The intrinsic binding constant, K =2.1 × 107M−1, of the complex to calf thymus DNA has been determined by absorption titration in 5 mmol dm−3 Tris-HCl, 50 mmol dm−3 NaCl buffer (pH 7.0). The excited state lifetimes and luminescence quenching with [Fe(CN)6]4− as the quencher in the presence of DNA were also tested and mono-exponentiality was observed for the emission decay curves. Viscosity measurements together with the optical titrations unambiguously proved that the complex bound with DNA intercalatively and that the binding affinity to DNA was several times larger than that of the parent complex [Ru(bpy)2(DPPZ)]2+.  相似文献   

8.
《BBA》1987,890(1):89-96
Electron donation to Photosystem II (PS II) by diphenylcarbazide (DPC) is interrupted by the presence of endogenous Mn in PS II particles. Removal of this Mn by Tris treatment greatly stimulates the electron transport with DPC as donor. Binding of low concentration of exogenous Mn(II) to Tris-treated PS II particles inhibits DPC photooxidation competitively with DPC. This phenomenon was used to locate a highly specific Mn(II) binding site on the oxidizing side of Photosystem II with dissociation constant about 0.15 μM. The binding of Mn(II) is electrostatic in nature. Its affinity depends not only on the ionic strength, but also on the anion species of the salt in the medium. The effectiveness in decreasing the affinity follows the order F > SO2−4 > CH3COO > CI > Br > NO3. This observation is interpreted as follows: smaller ions, like F, CH3COO, and larger ions, like SO2−4, have inhibitory effects on Mn(II) binding, whereas ions with optimal size, like Cl, Br and NO3, can stabilize the binding, resembling the anion requirement for reactivation of Cl-depleted chloroplasts. We suggest that the binding site for Mn(II) we observed is the site for the endogenous Mn in the O2-evolving complex of PS II. This site remains after Tris treatment, which removes all the endogenous Mn as well as the three extrinsic proteins, indicating that it is on the intrinsic component(s) of PS II reaction centers. Furthermore, the Cl requirement for O2 evolution may be attributed, at least partly to its stabilizing effect on Mn binding.  相似文献   

9.
Chloroform (CF) can undergo reductive dechlorination to dichloromethane, chloromethane, and methane. However, competition for hydrogen (H2), the electron-donor substrate, may cause poor dechlorination when multiple electron acceptors are present. Common acceptors in anaerobic environments are nitrate (NO3), sulfate (SO42−), and bicarbonate (HCO3). We evaluated CF dechlorination in the presence of HCO3 at 1.56 e Eq/m2-day, then NO3 at 0.04–0.15 e Eq/m2-day, and finally NO3 (0.04 e Eq/m2-day) along with SO42− at 0.33 e Eq/m2-day in an H2-based membrane biofilm reactor (MBfR). When the biofilm was initiated with CF-dechlorination conditions (no NO3 or SO42−), it yielded a CF flux of 0.14 e Eq/m2-day and acetate production via homoacetogenesis up to 0.26 e eq/m2-day. Subsequent addition of NO3 at 0.05 e Eq/m2-day maintained full CF dechlorination and homoacetogenesis, but NO3 input at 0.15 e Eq/m2-day caused CF to remain in the reactor's effluent and led to negligible acetate production. The addition of SO42− did not affect CF reduction, but SO42− reduction significantly altered the microbial community by introducing sulfate-reducing Desulfovibrio and more sulfur-oxidizing Arcobacter. Dechloromonas appeared to carry out CF dechlorination and denitrification, whereas Acetobacterium (homoacetogen) may have been involved with hydrolytic dechlorination. Modifications to the electron acceptors fed to the MBfR caused the microbial community to undergo changes in structure that reflected changes in the removal fluxes.  相似文献   

10.
In M. braunii, the uptake of NO3 and NO2 is blue-light-dependent and is associated with alkalinization of the medium. In unbuffered cell suspensions irradiated with red light under a CO2-free atmosphere, the pH started to rise 10s after the exposure to blue light. When the cellular NO3 and NO2 reductases were active, the pH increased to values of around 10, since the NH4+ generated was released to the medium. When the blue light was switched off, the pH stopped increasing within 60 to 90s and remained unchanged under background red illumination. Titration with H2SO4 of NO3 or NO2 uptake and reduction showed that two protons were consumed for every one NH4+ released. The uptake of Cl was also triggered by blue light with a similar 10 s time response. However, the Cl -dependent alkalinization ceased after about 3 min of blue light irradiation. When the blue light was turned off, the pH immediately (15 to 30 s) started to decline to the pre-adjusted value, indicating that the protons (and presumably the Cl) taken up by the cells were released to the medium. When the cells lacked NO3 and NO2 reductases, the shape of the alkalinization traces in the presence of NO3 and NO2 was similar to that in the presence of Cl, suggesting that NO3 or NO2 was also released to the medium. Both the NO3 and Cl-dependent rates of alkalinization were independent of mono- and divalent cations.  相似文献   

11.
Characterization of a proton pump from pea stem microsomes   总被引:1,自引:1,他引:0  
Abstract The present work deals with the characterization of an ATP-dependent proton translocation monitored by the ΔpH probe acridine orange. The ATP-dependent proton translocation has an optimum activity at pH 6.5 and is substrate specific for ATP. It is stimulated by Cl, HCO3 and Br, but is insensitive to several monovalent cations. Divalent cations (Mg2+ or Mn2+) are required for proton translocation, while in the presence of Ca2+ no uptake is observed. NO3, NO2 and citrate strongly inhibit proton uptake. On the contrary, F, SO42−, malate, pyruvate, succinate, oxalate and acetate have no inhibitory effect. Proton uptake is stimulated by valinomycin and unaffected by molybdate. Two thiols, dithioerythritol and dithiothreitol, are able partially to prevent the FCCP-abolished proton uptake or partially restore the ATP-dependent proton translocation in FCCP-collapsed vesicles. It is suggested that pea stem microsomes possess an electrogenic ATPase, acting as a proton pump, which, on the basis of its characteristics, can be tentatively associated with membranes of tonoplast origin.  相似文献   

12.
《Inorganica chimica acta》2006,359(7):2285-2290
Stopped-flow kinetic measurements were used to compare the reactivities of [Ru(medtra)(H2O)] (medtra3− = N-methylethylenediaminetriacetate) (1) and [Ru(hedtra)(H2O)] (2) (hedtra3− = N-hydroxyethylethylenediaminetriacetate) with NO in aqueous solution at 15 °C, pH 7.2 (phosphate buffer). The measured second-order rate constants (3 × 103 and 6 × 104 M−1 s−1 for 1 and 2, respectively) are three to four order of magnitudes lower than that for the reaction between [RuIII(edta)(H2O)] (3) with NO. However, NO scavenging studies of complexes 13, conducted by measuring the difference in nitrite production between treated and untreated murine macrophage cells, revealed that despite being less kinetically reactive toward NO, the [Ru(medtra)(H2O)] complex exhibited the highest NO scavenging ability and lowest toxicity of compounds 13.  相似文献   

13.
《Inorganica chimica acta》1986,114(2):151-158
CoX2(NO)(PMe3)2 complexes (X = Cl, Br, I, NO2) exhibit markedly different ν(NO) stretching frequencies and different geometries. The structure of CoI2(NO)(PMe3)2 (1) and CoCl2(NO)(PMe3)2 (2) have been determined by X-ray diffraction. Both crystallize in the orthorhombic system, Pnma space group with four molecules in a cell of the following dimensions: for 1, a = 10.497(2), b = 10.694(2), c = 13.975(2) Å, ν= 1568.8, Å3; for 2, a = 9.607(2), b = 10.689(2), c = 13.512(3) Å, ν= 1387.5 Å3. The structures were refined to conventional R values of R = 0.040 from 1630 reflections for 1 and R = 0.033 from 976 reflections for 2. In both cases, the coordination geometry about the five-coordinate cobalt atom is approximately trigonal bipyramidal, with the NO group sharing the equatorial positions with the halide ligands. Structure 2 is disordered, which prevents any precise structural characterization. In (1), the CoNO angle is 179.2(19)° and the Co NO distance is 1.728(23) Å; v(NO) is 1753 cm−1. CoCl2(NO)(PMe3)3 shows a v(NO) vibration at 1637 cm−1. Co(NO2)2(NO)(PMe3)2 with v(NO) = 1658 cm−1 has been proposed as a square pyramidal structure with a bent apical CoNO. These differences in NO stretching frequencies and geometries are discussed.  相似文献   

14.
A global scale Dynamic Nitrogen scheme (DyN) has been developed and incorporated into the Lund–Posdam–Jena (LPJ) dynamic global vegetation model (DGVM). The DyN is a comprehensive process‐based model of the cycling of N through and within terrestrial ecosystems, with fully interactive coupling to vegetation and C dynamics. The model represents the uptake, allocation and turnover of N in plants, and soil N transformations including mineralization, N2 fixation, nitrification and denitrification, NH3 volatilization, N leaching, and N2, N2O and NO production and emission. Modelled global patterns of site‐scale nitrogen fluxes and reservoirs are highly correlated to observations reported from different biomes. The simulation of site‐scale net primary production and soil carbon content was improved relative to the original LPJ, which lacked an interactive N cycle, especially in the temporal and boreal regions. Annual N uptake by global natural vegetation was simulated as 1.084 Pg N yr−1, with lowest values <1 g N m−2 yr−1 (polar desert) and highest values in the range 24–36.5 g N m−2 yr−1 (tropical forests). Simulated global patterns of annual N uptake are consistent with previous model results by Melillo et al. The model estimates global total nitrogen storage potentials in vegetation (5.3 Pg N), litter (4.6 Pg N) and soil (≥67 Pg as organic N and 0.94 Pg as inorganic N). Simulated global patterns of soil N storage are consistent with the analysis by Post et al. although total simulated N storage is less. Deserts were simulated to store 460 Tg N (up to 0.262 kg N m−2) as NO3, contributing 80% of the global total NO3 inventory of 580 Tg N. This model result is in agreement with the findings of a large NO3 pool beneath deserts. Globally, inorganic soil N is a small reservoir, comprising only 1.6% of the global soil N content to 1.5 m soil depth, but the ratio has a very high spatial variability and in hot desert regions, inorganic NO3 is estimated to be the dominant form of stored N in the soil.  相似文献   

15.
The platinum(II) complexes of the formula [Pt(DCHEDA)X2], where DCHEDA is N,N′-dicyclohexylethylenediamine and X is CL, Br, I, 0.5C2O42− (oxalate), 0.5C3H2O42− (malonate), 0.5C9H4O62− (4-carboxyphthalate), 0.5S2O32− or 0.5SO42−, have been synthesized and characterized by UVVis, IR, and 1H NMR spectral techniques. All the above complexes are non-electrolytes in DMF/H2O, except the sulphate complex which becomes a 1:1 electrolyte after incubation for 24 h at 28 °C. The halide complexes were also studied by X-ray photoelectron spectroscopy and these data suggest that there is π-bonding from platinum to halide in these complexes. The oxalate, malonate and sulphate bind in their complexes as bidentate ligands to platinum through two oxygen atoms whereas the thiosulphate in its complex binds as a bidentate ligand to platinum through one oxygen atom and one sulphur atom.  相似文献   

16.
The Arctic tundra has been shown to be a potentially significant regional sink for methyl chloride (CH3Cl) and methyl bromide (CH3Br), although prior field studies were spatially and temporally limited, and did not include gross flux measurements. Here we compare net and gross CH3Cl and CH3Br fluxes in the northern coastal plain and continental interior. As expected, both regions were net sinks for CH3Cl and CH3Br. Gross uptake rates (−793 nmol CH3Cl m−2 day−1 and −20.3 nmol CH3Br m−2 day−1) were 20–240% greater than net fluxes, suggesting that the Arctic is an even greater sink than previously believed. Hydrology was the principal regulator of methyl halide flux, with an overall trend towards increasing methyl halide uptake with decreasing soil moisture. Water table depth was one of the best predictors of net and gross uptake, with uptake increasing proportionately with water table depth. In drier areas, gross uptake was very high, averaging −1201 nmol CH3Cl m−2 day−1 and −34.9 nmol CH3Br m−2 day−1; in flooded areas, gross uptake was significantly lower, averaging −61 nmol CH3Cl m−2 day−1 and −2.3 nmol CH3Br m−2 day−1. Net and gross uptake was greater in the continental interior than in the northern coastal plain, presumably due to drier inland conditions. Within certain microtopographic features (low‐ and high‐centered polygons), uptake rates were positively correlated with soil temperature, indicating that temperature played a secondary role in methyl halide uptake. Incubations suggested that the inverse relationship between water content and methyl halide uptake was the result of mass transfer limitation in saturated soils, rather than because of reduced microbial activity under anaerobic conditions. These findings have potential regional significance, as the Arctic is expected to become warmer and drier due to anthropogenic climate forcing, potentially enhancing the Arctic sink for CH3Cl and CH3Br.  相似文献   

17.
The reaction of 8-thioguanosine (8-thioGuoH2 with methylmercury(II) has been shown to give rise to 1:1 (neutral and cationic), 1:2 (neutral and cationic), and 1:3 (cationic) complexes of the type [CH3Hg(8-thioGuoH)], [(CH3Hg(8-thioGuoH2)]NO3, [(CH3Hg)2(8-thioGuo)], [(CH3Hg)2(8-thioGuoH)]NO3 and [(CH3Hg)3(8-thioGuo)]NO3, depending upon the reactant stoichiometry and pH. 1H NMR, 13C NMR, and IR, as well as analytical data were used to characterize the complexes. Coupling of methylmercury(II)-protons to mercury-199 has been observed in all compounds. The magnitude of the coupling, 2J(1H-199Hg), is strongly dependent on the nature of the ligand bonded to the methylmercury(II) group and correlates with the 13C chemical shifts of coordinated CH3Hg(II) at the different binding sites.  相似文献   

18.
In an effort to better understand the biological efficacy of the tridentate aroyl hydrazone Cu(II) complexes, three Cu(II) complexes of acetylpyridine benzoyl hydrazone (HL), [Cu(L)(NO3) (H2O)]·H2O (C1), [Cu(L)2] (C2) and [Cu(L)(HL)]·(NO3)(Sas) (C3) (Sas = salicylic acid) were synthesized and characterized. X-ray crystal structures and infrared (IR) spectra of the complexes reveal that the L ligand of C1 and C2 are predominantly in the enolate resonance form, while one L ligand in C3 is represented enolate resonance form and the other HL ligand exhibits keto resonance form. All Cu(II) complexes showed significantly more anticancer activity than the ligand alone. Interestingly, the Cu complexes where the ligand/metal ratio was 1:1 (C1) rather than 2:1 (C2 and C3) had higher antitumor efficacy. Moreover, the 1:1 Cu/ligand complex, C1, promotes A549 cell apoptosis possibly through the intrinsic reactive oxygen species (ROS) mediated mitochondrial pathway, accompanied by the regulation of Bcl-2 family proteins.  相似文献   

19.
The green-tide macroalga, Ulva prolifera, was tested in the laboratory to determine its nutrient uptake and photosynthesis under different conditions. In the nutrient concentration experiments U. prolifera showed a saturated uptake for nitrate but an escalating uptake in the tested range for phosphorus. Both N/P and NO3 ?/NH4 + ratios influenced nutrient uptake significantly (p?<?0.05) while the PSII quantum yield [Y(II)] (p?>?0.05) remained unaffected. The maximum N uptake rate (33.9?±?0.8 μmol g?1 DW h?1) and P uptake rate (11.1?±?4.7) was detected at N/P ratios of 7.5 and 2.2, respectively. U. prolifera preferred NH4 +-N to NO3 ?-N when the NO3 ?-N/NH4 +-N ratio was less than 2.2 (p?<?0.05). But between ratios of 2.2 and 12.9, the uptake of NO3 ?-N surpassed that of NH4 +-N. In the temperature experiments, the highest N uptake rate and [Y(II)] were observed at 20 °C, while the lowest rates were detected at 5 °C. P uptake rates were correlated with increasing temperature.  相似文献   

20.
Investigation of side effects and solubility of anticancer drugs is a major challenge in chemotherapy science. Thus, design and synthesis of cisplatin analogs with higher lipophilicity as novel water-soluble anticancer drugs is valuable. In this work, two new Pt(II) complexes were synthesized with formula cis-[Pt(NH3)2(amylgly)]NO3 and cis-[Pt(amylamine)2(amylgly)]NO3, where gly is penthyl glycine as an amino acid. The new compounds were synthesized and extensively characterized using analytical techniques; spectroscopic methods, and conductivity measurement. The anticancer activity of synthesized complexes was investigated against colon cancer cell line HCT116 using MTT assay and results showed excellent anticancer activity with Cc50 values of 36 and 270 M after 24-h incubation time for cis-[Pt(NH3)2(amylgly)]NO3 and cis-[Pt(NH2-amyl)2(amylgly)]NO3, respectively; which is lower than that for cisplatin. These complexes were also interacted with highly polymerized calf thymus DNA and the binding mode of the complexes to CT-DNA was evaluated by fluorescence, circular dichroism, and UV spectroscopy. The calculation of binding and thermodynamic of Pt(II) complexes with CT-DNA can provide deeper insight into mechanism of the action of these types of complexes with nucleic acids. So, thermodynamic parameters were also determined according to isothermal titration. In comparison with cis-[Pt(NH3)2(amylgly)]NO3 in DNA interaction, the result show that cis-[Pt(NH2-amyl)2(amylgly)]NO3 has higher affinity with binding constant Kf = 8.72 mM to CT-DNA. The results indicate that cis-[Pt(amylamine)2(amylgly)]NO3 with large and bulky aliphatic group bind to CT-DNA by different modes and covalent and groove bindings were preferred mode of interaction with DNA.  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