首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The interaction of the VO2+ cation with the monoisoamyl ester of meso-2,3-dimercaptosuccinic acid (MiADMSA) was investigated by electron absorption spectroscopy in aqueous solutions at different pH values. The spectral behavior, complemented with a spectrophotometric titration, shows the generation of a [VO(MiADMSA)2]4− complex in which the oxocation interacts with two pairs of deprotonated –SH groups of the ester. Besides, MiADMSA rapidly reduces VO3 to VO2+, which might be chelated by an excess of the ester, and also produces relatively rapid reduction of V2O5 suspensions at pH = 6.5. The results of this study suggest that MiADMSA might be a potentially useful detoxification agent for vanadium.  相似文献   

2.
The interaction of the VO2+ cation with meso-2,3-dimercaptosuccinic acid (DMSA) was investigated by electron absorption spectroscopy in aqueous solution at different pH values. The spectral behavior, complemented with a spectrophotometric titration, shows the generation of a [VO(DMSA)2]2− complex in which the oxocation interacts with two pairs of deprotonated-SH groups of the acid. It was also found that DMSA rapidly reduces VO3 to VO2+, which might be chelated by an excess of the acid. DMSA can also produce the partial reduction of a V2O5 suspension at pH=5.2. The results of this study suggest that DMSA might be a potentially useful detoxification agent for vanadium.  相似文献   

3.
 The apo protein of imidazole glycerol phosphate dehydratase (IGPD) from Saccharomyces cerevisiae combines stoichiometrically with certain specific divalent metal cations to assemble the catalytically active form comprising 24 protein subunits and tightly bound metal. VO2+ ions react similarly but, uniquely, result in a metallo-protein (VO-IGPD) with neither catalytic activity nor the ability to bind to the reaction intermediate analogue, 2-hydroxy-3-(1,2,4-triazol-1-yl) propylphosphonate. Since VO2+ apparently assembles the quaternary structure correctly, it is used in the present study as a spin probe to investigate the metal centre coordination environment by electron paramagnetic resonance (EPR) and electron nuclear double resonance (ENDOR) spectroscopy. At neutral pH, the EPR spectrum of VO-IGPD reveals at least three distinct VO2+ sub-spectra with one predominant at low pH. The spin Hamiltonian parameters for some of the sub-spectra are consistent with 51V having nitrogen in the inner-sphere equatorial coordination environment from, most probably, multiple coordinating histidines. Further evidence for inner-sphere nitrogen ligands is obtained from ENDOR spectroscopy. The spectra of the low rf region show signals from interactions with 14N which are consistent with couplings to the imino nitrogen of coordinated histidine residues. In addition a number of proton ENDOR line pairs are resolved. Of the few that disappear upon exchange of the protein into D2O, one most likely originates from the exchangeable proton of the N-H group of a coordinated histidine imidazole. 1H-ENDOR line pairs from non-exchangeable protons with splittings of approximately 3 MHz can be attributed to imidazole carbon protons. Thus, most of the couplings observed by ENDOR are consistent with being from the imidazole heterocycle of one or more histidine ligands. Received: 27 June 1996 / Accepted: 14 March 1997  相似文献   

4.
The interaction of the VO2+ cation with homocysteine, was investigated by electron absorption spectroscopy in aqueous solution at different metal-to-ligand ratios. The direct reduction of vanadate(V) solutions with homocysteine was also investigated. The results suggest that the interaction is different from that found in the case of cysteine and occurs through pairs of amino and carboxylate groups of the amino acid. The interaction of VO2+ with homocystine, the oxidation product of homocysteine, was also analyzed. The interest of the results in relation to vanadium metabolism and detoxification is briefly discussed.  相似文献   

5.
High-intensity exercise results in reduced substrate levels and accumulation of metabolites in the skeletal muscle. The accumulation of these metabolites (e.g. ADP, Pi and H+) can have deleterious effects on skeletal muscle function and force generation, thus contributing to fatigue. Clearly this is a challenge to sport and exercise performance and, as such, any intervention capable of reducing the negative impact of these metabolites would be of use. Carnosine (β-alanyl-l-histidine) is a cytoplasmic dipeptide found in high concentrations in the skeletal muscle of both vertebrates and non-vertebrates and is formed by bonding histidine and β-alanine in a reaction catalysed by carnosine synthase. Due to the pKa of its imidazole ring (6.83) and its location within skeletal muscle, carnosine has a key role to play in intracellular pH buffering over the physiological pH range, although other physiological roles for carnosine have also been suggested. The concentration of histidine in muscle and plasma is high relative to its K m with muscle carnosine synthase, whereas β-alanine exists in low concentration in muscle and has a higher K m with muscle carnosine synthase, which indicates that it is the availability of β-alanine that is limiting to the synthesis of carnosine in skeletal muscle. Thus, the elevation of muscle carnosine concentrations through the dietary intake of carnosine, or chemically related dipeptides that release β-alanine on absorption, or supplementation with β-alanine directly could provide a method of increasing intracellular buffering capacity during exercise, which could provide a means of increasing high-intensity exercise capacity and performance. This paper reviews the available evidence relating to the effects of β-alanine supplementation on muscle carnosine synthesis and the subsequent effects on exercise performance. In addition, the effects of training, with or without β-alanine supplementation, on muscle carnosine concentrations are also reviewed.  相似文献   

6.
The interactions of VO2+ with phytate to form both soluble and insoluble complexes, have been studied by electronic absorption spectroscopy. A soluble 1∶1 VO2+: phytate complex is formed at pH <1. At higher pH-values insoluble complexes are produced. Two different solid complexes, obtained respectively at pH=2 and 4, were isolated and characterized. The maximal bonding ratio of VO2+: phytate was found to be 4, on the basis of a pH binding profile.  相似文献   

7.
 The reaction mechanism for the hydroxylation of benzene and monofluorobenzene, catalysed by a ferryl-oxo porphyrin cation radical complex (compound) is described by electronic structure calculations in local spin density approximation. The active site of the enzyme is modelled as a six-coordinated (Por+)Fe(IV)O a2u complex with imidazole or H3CS as the axial ligand. The substrates under study are benzene and fluorobenzene, with the site of attack in para, meta and ortho position with respect to F. Two reaction pathways are investigated, with direct oxygen attack leading to a tetrahedral intermediate and arene oxide formation as a primary reaction step. The calculations show that the arene oxide pathway is distinctly less probable, that hydroxylation by an H3CS–coordinated complex is energetically favoured compared with imidazole, and that the para position with respect to F is the preferred site for hydroxylation. A partial electron transfer from the substrate to the porphyrin during the reaction is obtained in all cases. The resulting charge distribution and spin density of the substrates reveal the transition state as a combination of a cation and a radical σ-adduct intermediate with slightly more radical character in the case of H3CS as axial ligand. A detailed analysis of the orbital interactions along the reaction pathway yields basically different mechanisms for the modes of substrate–porphyrin electron transfer and rupture of the Fe–O bond. In the imidazole-coordinated complex an antibonding π*(Fe–O) orbital is populated, whereas in the H3CS–coordinated system a shift of electron density occurs from the Fe–O bond region into the Fe–S bond. Received: 1 July 1995 / Accepted: 18 December 1995  相似文献   

8.
Two galactomannans, GALMAN-A and GALMAN-B, were isolated from seeds of Mimosa scabrella (bracatinga), with deactivation and exposure to native enzymes, respectively. They were treated with oxovanadium(IV) and oxovanadium(V), designated (VO2+/VO3+) to form GALMAN-A:VO2+/VO3+ and GALMAN-B:VO2+/VO3+ complexes, respectively. The potentiometric studies provided the binding constants for the complexes and the resulting complexed species were a function of pH. 51V NMR spectra of GALMAN-A:VO2+/VO3+ and GALMAN-B:VO2+/VO3+ at pH 7.8 and at 30 °C indicated the occurrence of two types of complexes formed by oxovanadium ions and galactomannans. GALMAN-A:VO2+/VO3+ and GALMAN-B:VO2+/VO3+ caused loss of HeLa cells viability at concentrations of 50-200 μg/mL. GALMAN-A:VO2+/VO3+ exhibited low toxicity for 24 h, although GALMAN-B:VO2+/VO3+ was extremely toxic, since 50 μg/mL was sufficient to decrease HeLa cell viability after 48 h by 60%. GALMAN-A gave rise to a slight increase in cell proliferation after 48 h at 100 μg/mL, whereas GALMAN-B promoted a slight decrease at concentrations of 50-100 μg/mL. GALMAN-A:VO2+/VO3+ and GALMAN-B:VO2+/VO3+ exhibited a significant decrease in cell proliferation after 48 h, each reaching 60% inhibition at 5-10 μg/mL. The complexes which caused this effect were at concentrations 10 times lower than the uncomplexed polymers.  相似文献   

9.
The complexation of VO2+ ion with the high molecular mass components of the blood serum, human serum transferrin (hTf) and albumin (HSA), has been re-examined using EPR spectroscopy. In the case of transferrin, the results confirm those previously obtained, showing that VO2+ ion occupies three different binding sites, A, B1 and B2, distinguishable in the X-band anisotropic spectrum recorded in D2O. With albumin the results show that a dinuclear complex (VO)2dHSA is formed in equimolar aqueous solutions or with an excess of protein; in the presence of an excess of VO2+, the multinuclear complex (VO)xmHSA is the prevalent species, where x = 5-6 indicates the equivalents of metal ion coordinated by HSA. The structure of the dinuclear species is discussed and the donor atoms involved in the metal coordination are proposed on the basis of the measured EPR parameters. Two different binding modes of albumin can be distinguished varying the pH, with only one species being present at the physiological value. The results show that the previously named “strong” site is not the N-terminal copper binding site, and some hypothesis on the metal coordination is discussed, with the 51V Az values for the proposed donor sets obtained by DFT (density functional theory) calculations. Finally, preliminary results obtained in the ternary system VO2+/hTf/HSA are shown in order to determine the different binding strength of the two proteins. Due to the low VO2+ concentration used, the recording of the EPR spectra through the repeated acquisition of the weak signals is essential to obtain a good signal to noise ratio in these systems.  相似文献   

10.
The pH- and time-dependent reaction of [Pt(dien)(H2O)]2+ with the methionine- and histidine-containing peptides H-His-Gly-Met-OH and Ac-His-Ala-Ala-Ala-Met-NHPh at 313 K has been investigated by HPLC and NMR spectroscopy. For both peptides, initial relatively rapid formation of the kinetically favoured methionine S-bound complex is followed by slow intramolecular migration of the [Pt(dien)]2+ fragment to imidazole Nε 2 (or, in the case of H-His-Gly-Met-OH, to a much lesser extent to the competing imidazole Nδ 1) of the histidine side chain over a period of 500 h. Time-dependent studies for the pentapeptide at pH 8.0 demonstrate that this isomerization can take place by either direct S→Nε 2 migration or by a two-step mechanism involving initial Nε 2 coordination of a second [Pt(dien)]2+ fragment and subsequent cleavage of the orginal Pt-S bond in the resulting dinuclear complex. The rate of κSN ε 2 isomerization is markedly reduced on lowering the pH to 5.1. Received: 26 February 1999 / Accepted: 14 April 1999  相似文献   

11.
1. (1) VO3 combines with high affinity to the Ca2+-ATPase and fully inhibits Ca2+-ATPase and Ca2+-phosphatase activities. Inhibition is associated with a parallel decrease in the steady-state level of the Ca2+-dependent phosphoenzyme.
2. (2) VO3 blocks hydrolysis of ATP at the catalytic site. The sites for VO3 also exhibit negative interactions in affinity with the regulatory sites for ATP of the Ca2+-ATPase.
3. (3) The sites for VO3 show positive interactions in affinity with sites for Mg2+ and K+. This accounts for the dependence on Mg2+ and K+ of the inhibition by VO3. Although, with less effectiveness, Na+ substitutes for K+ whereas Li+ does not. The apparent affinities for Mg2+ and K+ for inhibition by VO3 seem to be less than those for activation of the Ca2+-ATPase.
4. (4) Inhibition by VO3 is independent of Ca2+ at concentrations up to 50 μM. Higher concentrations of Ca2+ lead to a progressive release of the inhibitory effect of VO3.
Keywords: Ca2+-ATPase; Vanadate inhibition; K+; Li+; (Red cell membrane)  相似文献   

12.
DFT (B3LYP and M06L) as well as ab initio (MP2) methods with Dunning cc-pVnZ (n?=?2,3) basis sets are employed for the study of the binding ability of the new class of protease inhibitors, i.e., silanediols, in comparison to the well-known and well-studied class of inhibitors with hydroxamic functionality (HAM). Active sites of metalloproteases are modeled by [R3M-OH2]2+ complexes, where R stands for ammonia or imidazole molecules and M is a divalent cation, namely zinc, iron or nickel (in their different spin states). The inhibiting activity is estimated by calculating Gibbs free energies of the water displacement by metal binding groups (MBGs) according to: [R3M-OH2]2+ + MBG → [R3M-MBG]2+ + H2O. The binding energy of silanediol is only a few kcal mol?1 inferior to that of HAM for zinc and iron complexes and is even slightly higher for the triplet state of the (NH3)3Ni2+ complex. For both MBGs studied in the ammonia model the binding ability is nearly the same, i.e., Fe2+(t) > Ni2+(t) > Fe2+(q) > Ni2+(s) > Zn2+. However, for the imidazole model the order is slightly different, i.e., Ni2+(t) > Fe2+(t) > Fe2+(q) > Ni2+(s) ≥ Zn2+. Equilibrium structures of the R3Zn 2+ complexes with both HAM and silanediol are characterized by the monodentate binding, but the bidentate character of binding increases on going to iron and nickel complexes. Two types of intermediates of the water displacement reactions for [(NH3)3M-OH2]2+ complexes were found which differ by the direction of the attack of the MBG. Hexacoordinated complexes exhibit bidentate bonding of MBGs and are lower in energy for M=Ni and Fe. For Zn penta- and hexacoordinated complexes have nearly the same energy. Intermediate complexes with imidazole ligands have only octahedral structures with bidentate bonding of both HAM and dimethylsilanediol molecules.  相似文献   

13.
Poly(l-histidine) and imidazole in the presence of copper cations have been investigated by means of Fourier transform infrared (IR) spectroscopy in the mid- and far-IR spectral range to establish specific marker bands of the copper-coordination site in metalloproteins as a function of pH as well as the effect of the coordination on the amino acid contributions. Whereas the well-known mid-IR region was specific for the secondary structure of the protein mimics, the far-IR region included contributions from the metal–ligand vibrations. The addition of copper led to secondary structure changes of poly(l-histidine) at neutral and basic pD and to specific shifts of ring vibrations. At pD 9.5 the addition of copper deprotonated the nitrogen atoms of the imidazole ring and the backbone. At neutral pD the copper cations were coordinated by the N3 atom of the imidazole ring. Copper–imidazole vibrations at neutral pD were observed at 154 and 128 cm−1. Signals observed at 313 and 162 cm−1 were assigned to metal–ligand vibrations arising from copper–poly(l-histidine) complexes with a negatively charged imidazole ring.  相似文献   

14.
The steric and charge requirements for binding of l-carnosine (β-alanyl-l-histidine) by bovine serum albumin were investigated with proton magnetic resonance (1HMR) spectrometry. The histidinyl side chain of the dipeptide is responsible for primary recognition by the binding site. Furthermore, recognition is specific to a particular orientation of the histidinyl side chain that is determined by the other amino acid residue of the dipeptide. It was found that, although salts do not have a great effect on the binding of carnosine to bovine serum albumin, this binding cannot be measured by equilibrium dialysis in the presence of salt because of formation of a complex Donnan equilibrium. Carnosine, which has been postulated to have a role in olfaction, binds to the crude particulate fraction of nasal olfactory epithelium in the same steric orientation as it does to bovine serum albumin. Therefore, we have used the binding of carnosine to bovine serum albumin as a model system to test potential competitive inhibitors of carnosine binding that ultimately could be tested for activity in the olfactory pathway. It was found that the binding of carnosine to bovine serum albumin is a good model of nonspecific binding of carnosine to tissue preparations but not of the specific binding of carnosine to the nasal olfactory epithelium. In addition to requiring the proper conformation of the histidinyl residue, the binding to olfactory epithelium also appears to require recognition of the β-alanyl residue and of substituents on the imidazole ring. Evidence is provided that the carnosine binding by the nasal olfactory epithelium demonstrated by 1HMR spectroscopy does not occur with the mature olfactory receptor neurons.  相似文献   

15.
Summary. Glutathione (reduced form GSH and oxidized form GSSG) constitutes an important defense against oxidative stress in the brain, and taurine is an inhibitory neuromodulator particularly in the developing brain. The effects of GSH and GSSG and glycylglycine, γ-glutamylcysteine, cysteinylglycine, glycine and cysteine on the release of [3H]taurine evoked by K+-depolarization or the ionotropic glutamate receptor agonists glutamate, kainate, 2-amino-3-hydroxy-5-methyl-4-isoxazolepropionate (AMPA) and N-methyl-D-aspartate (NMDA) were now studied in slices from the hippocampi from 7-day-old mouse pups in a perfusion system. All stimulatory agents (50 mM K+, 1 mM glutamate, 0.1 mM kainate, 0.1 mM AMPA and 0.1 mM NMDA) evoked taurine release in a receptor-mediated manner. Both GSH and GSSG significantly inhibited the release evoked by 50 mM K+. The release induced by AMPA and glutamate was also inhibited, while the kainate-evoked release was significantly activated by both GSH and GSSG. The NMDA-evoked release proved the most sensitive to modulation: L-Cysteine and glycine enhanced the release in a concentration-dependent manner, whereas GSH and GSSG were inhibitory at low (0.1 mM) but not at higher (1 or 10 mM) concentrations. The release evoked by 0.1 mM AMPA was inhibited by γ-glutamylcysteine and cysteinylglycine, whereas glycylglycine had no effect. The 0.1 mM NMDA-evoked release was inhibited by glycylglycine and γ-glutamylcysteine. In turn, cysteinylglycine inhibited the NMDA-evoked release at 0.1 mM, but was inactive at 1 mM. Glutathione exhibited both enhancing and attenuating effects on taurine release, depending on the glutathione concentration and on the agonist used. Both glutathione and taurine act as endogenous neuroprotective effectors during early postnatal life. Authors’ address: Prof. Simo S. Oja, Brain Research Center, Medical School, FI-33014 University of Tampere, Finland  相似文献   

16.
Imidazole dipeptides, such as carnosine (β‐alanyl‐l ‐histidine) and anserine (β‐alanyl‐Nπ‐methyl‐l ‐histidine), are highly localized in excitable tissues, including skeletal muscle and nervous tissue, and play important roles such as scavenging reactive oxygen species and quenching reactive aldehydes. We have demonstrated several reactions between imidazole dipeptides (namely, carnosine, and anserine) and a lipid peroxide‐derived reactive aldehyde 4‐oxo‐2(E)‐nonenal. Seven carnosine adducts and two anserine adducts were characterized using liquid chromatography/electrospray ionization‐multiple‐stage mass spectrometry. Adduct formation occurred between imidazole dipeptides and 4‐oxo‐2(E)‐nonenal mainly through Michael addition, Schiff base formation, and/or Paal‐Knorr reaction. The reactions were much more complicated than the reaction with a similar lipid peroxide‐derived reactive aldehyde, 4‐hydroxy‐2(E)‐nonenal.  相似文献   

17.
D. Curtin  G. Wen 《Plant and Soil》2004,267(1-2):109-115
Plants that remove an excess of cations over anions may cause soil acidification. The acidification potential of plants has been evaluated using solution culture techniques, but the influence of ionic composition of the medium on the plant cation-anion balance remains unclear. Our objective was to determine how electrolyte concentration and salt type affect the cation- anion balance of two test plants [barley (Hordeum vulgare L.) and kochia (Kochia scoparia L. Schrad.)]. Seedlings were grown in sand culture and irrigated with nutrient solution (Hoagland’s solution), which was adjusted to a range of electrolyte concentrations (target electrical conductivity of 7.5, 17.5 and 27.5 dS m−1) using either chloride or sulphate salts. Increase in electrolyte concentration reduced yield of kochia, a salt-tolerant plant, by up to 38%. Total cation (Ca + Mg + K + Na) equivalents in kochia exceeded those of anions (Cl + S + P + NO3) by 250 to 280 cmolc kg−1 of dry matter. Electrolyte concentration had no effect on the cation-anion balance of kochia, but excess cation values were significantly greater in the sulphate than in the chloride system. Kochia had a large content of water-soluble oxalate (194 to 226 cmolc kg−1), which was linearly related to the excess cation content. Growth of barley was severely restricted at the intermediate and high electrolyte concentrations. Cations exceeded anions by 21 to 59 cmolc kg−1 of barley dry matter. Excess cation content was greater in the sulphate than in the chloride medium, but electrolyte concentration did not have a consistent effect on the cation-anion balance. The small amounts of oxalate found in barley (0.9 to 2.6 cmolc kg−1) were insufficient to balance the cation excess.  相似文献   

18.
The proligands PicMe-AaR (PicMe = methoxipicolyl-5-amide, where the amide substituent is an amino acid AaR = HisH, HisMe, IleH, IleMe, TrpH, TrpMe, HTyrEt, tBuTyrMe, HThrMe, tBuThrMe) and the complexes [VO(Pic-AaR)2] have been synthesised and characterised. A detailed EPR study of the VO2+/Pic-His systems in water revealed the predominance of the complex [VO(Pic-His)H2O] in the pH range 2-6, with tridentate coordination of Pic-His via the picolinate moiety and imidazole-Nδ. Speciation analyses of the binary systems VO2+/Pic-Aa (Aa = His, Ile, Trp) and the ternary systems VO2+/Pic-Aa/B (Aa = His, Ile; B = citrate (cit), lactate (lac), phosphate) showed a predominance of the ternary complexes [VO(Pic-Aa)(cit/lac)] and [VO(Pic-Aa)(cit/lac)OH] in the physiological pH regime. If, in addition, human serum albumin (HAS) and apotransferrin (Tf) are present, with all of the low and high molecular mass constituents in their blood serum concentrations, about two thirds of VO2+ is bound to the protein, while there is still a sizable amount of ternary complex [VO(Pic-Aa)(cit/lac)] present (about 1/4 for Pic-His and 1/3 for Pic-Ile) when the vanadium(IV) concentration is relatively high; at lower concentrations Tf is the predominant binder. Insulin-mimetic studies for VO2+/Pic-Aa (Aa = His, Ile, Tyr and Trp), based on a lipolysis assay with rat adipocytes, provided IC50 values of 0.41(1) for VO2+/Pic-His and VO2+/Pic-Ile, which compares with 0.87(17) for VOSO4.  相似文献   

19.
Cobalt(II) ion and L-carnosine produce two different complexes when mixed in aqueous solution at pH 7.2. One complex has coordination of N-3 of the imidazole ring to the cobalt(II) and is produced when the concentration of peptide exceeds that of cobalt(II). The second complex has chelation of three nitrogen atoms of a single carnosine. This second complex produces a reversible oxygen carrier by making stable mixed chelates with additional carnosine, histidine or cysteine. These results indicate that cobalt complexes with mixed ligands should be of more importance invivo than those with carnosine as the only ligand. They provide an explanation for the high activity and substrate specificity of carnosinase in kidney.  相似文献   

20.
Store‐operated calcium entry (SOCE) is essential for many cellular processes. In this study, we investigated modulation of SOCE by tyrosine phosphorylation in rat epididymal basal cells. The intracellular Ca2+([Ca2+]i) measurement showed that SOCE occurred in rat epididymal basal cells by pretreating the cells with thapsigargin (Tg), the inhibitor of sarco‐endoplasmic reticulum Ca2+‐ATPase. To identify the role of Ca2+ channels in this response, we examined the effects of transient receptor potential canonical channel blockers 2‐aminoethoxydiphenyl borate (2‐APB), 1‐[β‐[3‐(4‐methoxyphenyl)pro‐poxy]‐4‐methoxyphenethyl]‐1H‐imidazole hydrochloride(SKF96365), Gd3+, and non‐selective cation channel blocker Ni2+ respectively on SOCE and found that these blockers could inhibit the Ca2+ influx to different extent. Furthermore, we studied the regulation of SOCE by tyrosine kinase pathway. The inhibitor of tyrosine kinase genistein remarkably suppressed the SOCE response, whereas sodium orthovanadate, the inhibitor of tyrosine phosphatase, greatly enhanced it. The results suggest that tyrosine kinase pathway plays a significant role in the initiation of SOCE and positively modulates SOCE in epididymal basal cells. J. Cell. Physiol. 226: 1069–1073, 2011. © 2010 Wiley‐Liss, Inc.  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