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.
Elevation of nitric oxide (NO˙) can either promote or inhibit ultraviolet B light (UVB)-induced apoptosis. In this study, we determined real-time concentration of NO˙ and peroxynitrite (ONOO) and their role in regulation of membrane integrity and apoptosis. Nanosensors (diameter 300–500 nm) were used for direct in situ simultaneous measurements of NO˙ and ONOO generated by UVB in cultured keratinocytes and mice epidermis. An exposure of keratinocytes to UVB immediately generated ONOO at maximal concentration of 190 nm followed by NO˙ release with a maximal concentration of 91 nm . The kinetics of UVB-induced NO˙/ONOO was in contrast to cNOS agonist stimulated NO˙/ONOO from keratinocytes. After stimulating cNOS by calcium ionophore (CaI), NO˙ release from keratinocytes was followed by ONOO production. The [NO˙] to [ONOO] ratio generated by UVB decreased below 0.5 indicating a serious imbalance between cytoprotective NO˙ and cytotoxic ONOO—a main component of nitroxidative stress. The NO˙/ONOO imbalance increased membrane damage and cell apoptosis was partially reversed in the presence of free radical scavenger. The results suggest that UVB-induced and cNOS-produced NO˙ is rapidly scavenged by photolytically and enzymatically generated superoxide (O2˙) to produce high levels of ONOO, which enhances oxidative injury and apoptosis of the irradiated cells.  相似文献   

2.
This article presents a theoretical investigation of the reaction mechanism of imidazole nitration by peroxynitrite using density functional theory calculations. Understanding this reaction mechanism will help in elucidating the mechanism of guanine nitration by peroxynitrite, which is one of the assumed chemical pathways for damaging DNA in cells. This work focuses on the analysis of the potential energy surface (PES) for this reaction in the gas phase. Calculations were carried out using Hartree–Fock (HF) and density functional theory (DFT) Hamiltonians with double‐zeta basis sets ranging from 6‐31G(d) to 6‐31++G(d,p), and the triple‐zeta basis set 6‐311G(d). The computational results reveal that the reaction of imidazole with peroxynitrite in gas phase produces the following species: (i) hydroxide ion and 2‐nitroimidazole, (ii) hydrogen superoxide ion and 2‐nitrosoimidazole, and (iii) water and 2‐nitroimidazolide. The rate‐determining step is the formation of a short‐lived intermediate in which the imidazole C2 carbon is covalently bonded to peroxynitrite nitrogen. Three short‐lived intermediates were found in the reaction path. These intermediates are involved in a proton‐hopping transport from C2 carbon to the terminal oxygen of the ? O? O moiety of peroxynitrite via the nitroso (ON? ) oxygen. Both HF and DFT calculations (using the Becke3–Lee–Yang–Parr functional) lead to similar reaction paths for proton transport, but the landscape details of the PES for HF and DFT calculations differ. This investigation shows that the reaction of imidazole with peroxynitrite produces essentially the same types of products (nitro‐ and nitroso‐) as observed experimentally in the reaction of guanine with peroxynitrite, which makes the former reaction a good model to study by computation the essential characteristics of the latter reaction. Nevertheless, the computationally determined activation energy for imidazole nitration by peroxynitrite in the gas phase is 84.1 kcal/mol (calculated at the B3LYP/6‐31++G(d,p) level), too large for an enzymatic reaction. Exploratory calculations on imidazole nitration in solution, and on the reaction of 9‐methylguanine with peroxynitrite in the gas phase and solution, show that solvation increases the activation energy for both imidazole and guanine, and that the modest decrease (15 kcal mol?1) in the activation energy, due to the adjacent six member ring of guanine, is counterbalanced by solvation. These results lead to the speculation that proton tunneling may be at the origin of experimentally observed high reaction rate of guanine nitration by peroxynitrite in solution. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

3.
During the decay of (15N)peroxynitrite (O?15NOO ? ) in the presence of N‐acetyl‐L ‐tyrosine (Tyrac) in neutral solution and at 268 K, the 15N‐NMR signals of 15NO and 15NO show emission (E) and enhanced absorption (A) as it has already been observed by Butler and co‐workers in the presence of L ‐tyrosine (Tyr). The effects are built up in radical pairs [CO , 15NO ]S formed by O? O bond scission of the (15N)peroxynitrite? CO2 adduct (O?15NO? OCO ). In the absence of Tyrac and Tyr, the peroxynitrite decay rate is enhanced, and 15N‐CIDNP does not occur. This is explained by a chain reaction during the peroxynitrite decay involving N2O3 and radicals NO . and NO . The interpretation is supported by 15N‐CIDNP observed with (15N)peroxynitrite generated in situ during reaction of H2O2 with N‐acetyl‐N‐(15N)nitroso‐dl ‐tryptophan ((15N)NANT) at 298 K and pH 7.5. In the presence of Na15NO2 at pH 7.5 and in acidic solution, 15N‐CIDNP appears in the nitration products of Tyrac, 1‐(15N)nitro‐N‐acetyl‐L ‐tyrosine (1‐15NO2‐Tyrac) and 3‐(15N)nitro‐N‐acetyl‐L ‐tyrosine (3‐15NO2‐Tyrac). The effects are built up in radical pairs [Tyrac . , 15NO ]F formed by encounters of independently generated radicals Tyrac . and 15NO . Quantitative 15N‐CIDNP studies show that nitrogen dioxide dependent reactions are the main if not the only pathways for yielding both nitrate and nitrated products.  相似文献   

4.
生物体内NO和超氧阴离子快速反应生成的过氧亚硝酸根离子(ONOO-,peroxynitrite)是一种强细胞毒性物质,它诱导蛋白质酪氨酸残基硝化是其损伤生物系统的重要途径之一。为了探讨谷胱甘肽和ebselen对胰岛素硝化的抑制及其相互作用机理,采用UV-Vis、HPLC和ESI-MS等方法,研究了ONOO-对胰岛素的硝化作用、小分子抗氧化剂谷胱甘肽(GSH)和ebselen对ONOO-硝化胰岛素的影响以及它们之间的相互作用。结果表明单独的GSH和ebselen对ONOO--引发的胰岛素硝化均有明显的抑制,而作为谷胱甘肽过氧化物酶(GPx)的底物GSH 与GPx的模型化合物ebselen之间存在相互拮抗作用,经过对其产物分析,确定其机理是GSH和ebselen能够直接反应生成一种加合物,从而抑制了GSH和ebselen各自的抗硝化能力。  相似文献   

5.
Nitryl chloride and peroxynitrite are reactive nitrogen species generated by activated phagocytes against invading pathogens during infections and inflammation. In our previous report, formation of 8‐nitroxanthine and 8‐nitroguanine was observed in reaction of 2′‐deoxyguanosine or calf thymus DNA with nitryl chloride generated by mixing hypochlorous acid (HOCl) with nitrite (NC2?). The present study investigates factors control ling the yields of 8‐nitroxanthine and 8‐nitroguanine formation in nitration of 2′‐deoxyguanosine by nitryl chloride. We found that the yields of 8‐nitroxanthine and 8‐nitroguanine in reaction of 2′‐deoxyguanosine with nitryl chloride were highly dependent on the ratio of NO2? versus HOCl concentration. The yields of 8‐nitroxanthine and 8‐nitroguanine reached a plateau when the ratio of NC2? versus HOCl concentration was higher than 2. A possible mechanism was postulated to explain this observation. While 8‐nitroguanine is not stable in the presence of peroxynitrite, 8‐nitroxanthine is sensitive to HOCl. The stability of these two nitrated ad ducts might be a factor on their final yields in this reaction. Since HOCl is produced by neutrophils at sites of inflammation where the level of NC2? is elevated, it is conceivable that nitryl chloride contributes to DNA base nitration in vivo, forming 8‐nitroxanthine and 8‐nitroguanine.  相似文献   

6.
Nitrated fatty acids (NO2‐FAs) exhibit a variety of important biological attributes, including a nitric oxide (˙NO) donor and a cell‐signaling molecule. We investigated the mechanisms of fatty‐acid nitration, and the release of ˙NO from NO2‐FAs. NO2‐FAs are formed effectively by the addition of ˙NO2, followed by either hydrogen abstraction or addition of a second NO2. The latter reaction results in a vicinal nitronitrite ester form of FA, which isomerizes into vicinal nitrohydroxy FA via hydronium ion catalysis. The nitrohydroxy FAs exist in equilibria with NO2‐FAs. Nitration of conjugated linoleic acid (cLA) was proved to be significantly more efficient than that of LA. In a nonaqueous environment, release of ˙NO from nitrite ester (ONO‐FA) was facilitated by ˙NO2. Furthermore, the release of ˙NO from NO2‐cLA is the most favorable in the nitrite ester mechanism. In an aqueous environment, the modified Nef reaction was shown to be feasible. In addition, the release of ˙NO from 10‐ and 12‐NO2‐LA involves a larger reaction barrier and is more endergonic than those from 9‐ and 13‐NO2‐LA.  相似文献   

7.
Peroxynitrite (ONOO, oxoperoxonitrate(1−)), an isomer of nitrate that oxidizes and nitrates biomolecules, is likely to be formed in vivo from the reaction of superoxide with nitrogen monoxide. To determine whether flavonoids scavenge peroxynitrite as postulated in the literature, we studied the reactions of peroxynitrite with phenol, hydroquinone, catechol, and the flavonoid monoHER. These reactions are first‐order with respect to peroxynitrous acid and zero‐order with respect to the organic compounds, and proceed as fast as the isomerization of peroxynitrous acid to nitrate. In vivo, a large fraction of all peroxynitrite is likely to react with carbon dioxide to form an unstable adduct, the 1‐carboxylato‐2‐nitrosodioxidane anion (ONOOCO). The presence of phenolic compounds did not influence the rate of disappearance of this adduct, which was ca. 4×102 s−1. On the basis of these kinetic studies, it can be concluded that flavonoids are not scavengers of peroxynitrite. The products from the reaction of peroxynitrite with hydroquinone (benzene‐1,4‐diol) and catechol (benzene‐1,2‐diol) are para‐benzoquinone and ortho‐benzoquinone, respectively; no nitrated products were found. In a subsequent reaction, ortho‐quinone reacted with nitrite, a common contaminant of peroxynitrite preparations to form 1,2‐dihydroxy‐4‐nitrobenzene. We also investigated whether carbonyl compounds could redirect the reactivity of peroxynitrite toward nitration, as carbon dioxide does. The reaction with acetone is first‐order with respect to peroxynitrite and first‐order with respect to the carbonyl compound. The rate constant is 1.8 M −1s−1 at neutral pH and 20°; peroxynitrite does not react with the carbonyl compounds dimethyl acetamide, L ‐alanylalanine, or methyl acetate. It is not likely that the carbonyl compounds or the mono‐, di‐, or polyphenolic compounds can scavenge peroxynitrite in vivo.  相似文献   

8.
N‐(p‐Nitrophenoxy)carbonyl‐3‐morpholino‐sydnonimine (NCMS) has been prepared from 3‐morpholinosydnonimine hydrochloride. Using the Griess assay and the superoxide‐mediated reduction of ferricytochrome c, the nitric oxide (NO?) and superoxide anion (O2?) ‐ releasing properties in phosphate buffer pH 7.4 of this novel peroxynitrite donor was studied and compared with the known 3‐morpholino‐sydnonimine (SIN‐1). From compound NCMS, a series of N‐substituted sydnonimine derivatives were easily prepared that contain purine or melaminophenyl groups which specify a recognition by a trypanosomal purine transporter. The ability of these new sydnonimines to inhibit the uptake of [23H]adenosine on Trypanosoma equiperdum was studied.  相似文献   

9.
The mechanistic aspects of one‐electron oxidation of G‐quadruplexes in the basket (Na+ ions) and hybrid (K+ ions) conformations were investigated by transient absorption laser kinetic spectroscopy and HPLC detection of the 8‐oxo‐7,8‐dihydroguanine (8‐oxoG) oxidation product. The photo‐induced one‐electron abstraction from G‐quadruplexes was initiated by sulfate radical anions (SO4˙?) derived from the photolysis of persulfate ions by 308 nm excimer laser pulses. In neutral aqueous solutions (pH 7.0), the transient absorbance of neutral guanine radicals, G(‐H)˙, is observed following the complete decay of SO4˙? radicals (~10 μs after the actinic laser flash). In both basket and hybrid conformations, the G(‐H)˙ decay is biphasic with one component decaying with a lifetime of ~0.1 ms, and the other with a lifetime of 20–30 ms. The fast decay component (~0.1 ms) in G‐quadruplexes is correlated with the formation of 8‐oxoG lesions. We propose that in G‐quadruplexes, G(‐H)˙ radicals retain radical cation character by sharing the N1‐proton with the O6‐atom of G in the [G˙+: G] Hoogsteen base pair; this [G(‐H)˙: H+G G˙+: G] leads to the hydration of G˙+ radical cation within the millisecond time domain, and is followed by the formation of the 8‐oxoG lesions.  相似文献   

10.
It has been shown recently that photosystem 1 particles, photosystem 1 lipid vesicles and chlorophyll-a lipid vesicles show identical photochemical reactions in the presence of oxygen e.g. H+-and O2-uptake (Van Ginkel, 1979). Therefore, spin-trapping experiments were done to identify the oxygen radicals formed. The spintrap phenyltertiarybutylnitrone (PBN) failed to yield information about oxygen radicals. With the spintrap 5,5-dimethyl-1-pyrroline-1-oxide (DMPO), however, we obtained a mixed spectrum of O- and OH·-adducts generated in chloroplasts, photosystem 1 particles or chlorophyll-a lipid vesicles. These data indicate that chlorophyll-a in an artificial membrane can also catalyze O--formation. Chlorophyll-a lipid vesicles catalyze light-induced formation of the Tiron-semiquinone free radical, which has been proposed as a specific O--probe (Greenstock and Miller, 1975). However, OH· scavengers strongly reduce the formation of this radical, whereas superoxide dismutase does not. Pulse-radiolysis measurements showed that the rate constant for the reaction of Tiron with OH· is 8.2 · 109M-1 s-1, which is considerably higher than the published Tiron/O- rate constants. Therefore, Tiron is a better spin probe for OH· than for O-. We suggest that light-induced H+-and O--uptake in membranes containing chlorophyll-a in the presence of ascorbate is caused mainly by the very rapid reaction of OH· with ascorbate.  相似文献   

11.
The spin-trapping technique was used to study the radical intermediates produced by reaction of nitric oxide (*NO) and peroxynitrite with serum albumin and glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Our results show that the major radical product induced by *NO and by peroxynitrite with serum albumin and GAPDH was a thiyl radical. The same radical can be detected in the *NO-transfer from S-nitroso albumin to low molecular weight thiols. Moreover, *NO or peroxynitrite treatment of GAPDH was able to induce NAD-dependent covalent modification of the enzyme in erythrocyte ghosts.  相似文献   

12.
Tsikas D 《The Analyst》2011,136(5):979-987
Nitric oxide (˙NO) and superoxide (O(2)(-)˙) are ubiquitous in nature. Their reaction product peroxynitrite (ONOO(-)) and notably its conjugated peroxynitrous acid (ONOOH) are highly unstable in aqueous phase. ONOO(-)/ONOOH (referred to as peroxynitrite) isomerize and decompose to NO(3)(-), NO(2)(-) and O(2). Here, we report for the first time GC-MS and HPLC methods for the analysis of peroxynitrite in aqueous solution. For GC-MS analysis peroxynitrite in alkaline solution was derivatized to a pentafluorobenzyl derivative using pentafluorobenzyl bromide. O(15)NOO(-) was synthesized from H(2)O(2) and (15)NO(2)(-) and used as internal standard. HPLC analysis was performed on stationary phases consisting of Nucleosil? 100-5C(18)AB or Nucleodur? C(18) Gravity. The mobile phase consisted of a 10 mM aqueous solution of tetrabutylammonium hydrogen sulfate and had a pH value of 11.5. UV absorbance detection at 300 nm was used. HPLC allows simultaneous analysis of ONOO(-), NO(2)(-) and NO(3)(-). The GC-MS and HPLC methods were used to study stability, synthesis, formation from S-[(15)N]nitrosoglutathione (GS(15)NO) and KO(2), and isomerization/decomposition of peroxynitrite to NO(2)(-) and NO(3)(-) in aqueous buffer.  相似文献   

13.
The reaction of peroxynitrite/peroxynitrous acid with morpholine as a model compound for secondary amines is reinvestigated in the absence and presence of carbon dioxide. The concentration‐ and pH‐dependent formation of N‐nitrosomorpholine and N‐nitromorpholine as reported in three previous papers ([25] [26] [14]) is basically confirmed. However, 13C‐NMR spectroscopic product analysis shows that, in the absence of CO2, N‐hydroxymorpholine is, at pH ≥ 7, the major product of this reaction, even under anaerobic conditions. The formation of N‐hydroxymorpholine has been overlooked in the three cited papers. Additional (ring‐opened) oxidation products of morpholine are also detected. The data account for radical pathways for the formation of these products via intermediate morpholine‐derived aminyl and α‐aminoalkyl radicals. This is further supported by EPR‐spectrometric detection of morpholine‐derived nitroxide radicals, i.e., morpholin‐4‐yloxy radicals. N‐Nitrosomorpholine, however, is very likely formed by electrophilic attack of peroxynitrite‐derived N2O4. 15N‐CIDNP Experiments establish that, in the presence of CO2, N‐nitro‐ and C‐nitromorpholine are generated by radical recombination. The present results are in full accord with a fractional (28 ± 2%) homolytic decay of peroxynitrite/peroxynitrous acid with release of free hydroxyl and nitrogen dioxide radicals.  相似文献   

14.
The elementary mechanism of radical-mediated peptide tyrosine nitration, which is a hallmark of post-translational modification of proteins under nitrative stress in vivo, has been elucidated in detail by using an integrated approach that combines the gas-phase synthesis of prototypical molecular tyrosine-containing peptide radical cations, ion–molecule reactions, and isotopic labeling experiments with DFT calculations. This reaction first involves the radical recombination of .NO2 towards the prerequisite phenoxyl radical tautomer of a tyrosine residue, followed by proton rearrangements, finally yielding the stable and regioselective 3-nitrotyrosyl residue product. In contrast, nitration with the π-phenolic radical cation tautomer is inefficient. This first direct experimental evidence for the elementary steps of the radical-mediated tyrosine nitration mechanism in the gas phase provides a fundamental insight into the regioselectivity of biological tyrosine ortho-nitration.  相似文献   

15.
16.
Tyrosine nitration, often observed during neurodegenerative disorders under nitrative stress, is usually considered to be induced chemically either by nitric oxide and oxygen forming nitrogen dioxide or by the decomposition of peroxynitrite. It can also be induced enzymatically by peroxidases or superoxide dismutases in the presence of both hydrogen peroxide and nitrite forming nitrogen dioxide and/or peroxynitrite. In this study, the role of cupric ions for catalyzing tyrosine nitration in the presence of hydrogen peroxide and nitrite, by a chemical mechanism rather similar to enzymatic pathways where nitrite is oxidized to form nitrogen dioxide, was investigated by development of a microreactor also capable of acting as an emitter for electrospray ionization mass spectrometry analysis. Indeed, cupric ions and peptide-cupric ion complexes are found to be excellent Fenton catalysts, even better than Fe(III) or heme, for the formation of (?)OH radicals and/or copper(II)-bound (?)OH radicals from hydrogen peroxide. These radicals are efficiently scavenged by nitrite anions to form (?)NO(2) and by tyrosine to form tyrosine radicals, leading to tyrosine nitration in polypeptides. We also show that cupric ions can catalyze tyrosine nitration from nitric oxide, oxygen, and hydrogen peroxide as the formation of tyrosine radicals is increased in the presence of diffusible and/or copper(II) bound hydroxyl radicals. This study shows that copper has a polyvalent role in the processes of tyrosine nitration.  相似文献   

17.
A major barrier to understanding the mechanism of nitric oxide reductases (NORs) is the lack of a selective probe of NO binding to the nonheme FeB center. By replacing the heme in a biosynthetic model of NORs, which structurally and functionally mimics NORs, with isostructural ZnPP, the electronic structure and functional properties of the FeB nitrosyl complex was probed. This approach allowed observation of the first S=3/2 nonheme {FeNO}7 complex in a protein‐based model system of NOR. Detailed spectroscopic and computational studies show that the electronic state of the {FeNO}7 complex is best described as a high spin ferrous iron (S=2) antiferromagnetically coupled to an NO radical (S= 1/2) [Fe2+‐NO.]. The radical nature of the FeB‐bound NO would facilitate N? N bond formation by radical coupling with the heme‐bound NO. This finding, therefore, supports the proposed trans mechanism of NO reduction by NORs.  相似文献   

18.
We designed and synthesized a photocontrollable peroxynitrite (ONOO(-)) generator, P-NAP, which has N-methyl-N-nitrosoaminophenol structure with four methyl groups introduced onto the benzene ring to block reaction of the photodecomposition product with ONOO(-) and to lower the semiquinoneimine's redox potential. The semiquinoneimine intermediate generated by photoinduced release of nitric oxide (NO) reduces dissolved molecular oxygen to generate superoxide radical anion (O(2)(?-)), which reacts with NO to afford ONOO(-) under diffusion control (k = 6.7 × 10(9) M(-1) s(-1)). NO release from P-NAP under UV-A (330-380 nm) irradiation was confirmed by ESR spin trapping. Tyrosine nitration, characteristic of ONOO(-), was demonstrated by HPLC analysis of a photoirradiated aqueous solution of P-NAP and N-acetyl-l-tyrosine ethyl ester. ONOO(-) formation was confirmed with a ONOO(-)-specific fluorogenic probe, HKGreen-3, and compared with that from 3-(4-morpholinyl)sydnonimine hydrochloride (SIN-1), which is the most widely used ONOO(-) generator at present. The photoreaction of P-NAP was influenced by superoxide dismutase, indicating that generation of O(2)(?-) occurs before ONOO(-) formation. The quantum yield for formation of duroquinone, the main P-NAP photodecomposition product, was measured as 0.86 ± 0.07 at 334 nm with a potassium ferrioxalate actinometer. Generation of ONOO(-) from P-NAP in HCT-116 cells upon photoirradiation was successfully imaged with HKGreen-3A. This is the first example of a photocontrollable ONOO(-) donor applicable to cultured cells.  相似文献   

19.
Peroxynitrite (ONOOH/ONOO-) which is formed in vivo under oxidative stress is a strong oxidizing and nitrating agent. It has been reported that several flavonoids, including quercetin, inhibit the peroxynitrite-induced oxidation and/or nitration of several molecules tested; however, the mechanism of their protective action against peroxynitrite is not univocally resolved. The kinetics of the reaction of quercetin with peroxynitrite was studied by stopped-flow as well as by conventional spectrophotometry under acidic, neutral and alkaline pH. The obtained results show that the protective mechanism of quercetin against peroxynitrite toxicity cannot be explained by direct scavenging of peroxynitrite. We propose that quercetin acts via scavenging intermediate radical products of peroxynitrite decomposition (it is an excellent scavenger of ·NO2) and/or via reduction of target radicals formed in the reaction with peroxynitrite.  相似文献   

20.
A kinetic study of the one electron oxidation of substituted ferrocenes (FcX: X = H, COPh, COMe, CO(2)Et, CONH(2), CH(2)OH, Et, and Me(2)) by a series of N-oxyl radicals (succinimide-N-oxyl radical (SINO), maleimide-N-oxyl radical (MINO), 3-quinazolin-4-one-N-oxyl radical (QONO) and 3-benzotriazin-4-one-N-oxyl radical (BONO)), has been carried out in CH(3)CN. N-oxyl radicals were produced by hydrogen abstraction from the corresponding N-hydroxy derivatives by the cumyloxyl radical. With all systems, the rate constants exhibited a satisfactory fit to the Marcus equation allowing us to determine self-exchange reorganization energy values (λ(NO˙/NO(-))) which have been compared with those previously determined for the PINO/PINO(-) and BTNO/BTNO(-) couples. Even small modification of the structure of the N-oxyl radicals lead to significant variation of the λ(NO˙/NO(-)) values. The λ(NO˙/NO(-)) values increase in the order BONO < BTNO < QONO < PINO < SINO < MINO which do not parallel the order of the oxidation potentials. The higher λ(NO˙/NO(-)) values found for the MINO and SINO radicals might be in accordance with a lower degree of spin delocalization in the radicals MINO and SINO and charge delocalization in the anions MINO(-) and SINO(-) due to the absence of an aromatic ring in their structure.  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