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 93 毫秒
1.
The 95Mo NMR spectra of a series of seven-coordinate molybdenum(II) isocyanide complexes of the types [Mo(CNR)7-nLn](PF6)2 (R = CH3, CHMe2, CMe3, C6H11, CH2Ph; L = py, bpy, Me2bpy, phen, dppe, P-n-Bu3; n = 0,1,2) [Mo(CNC-Me3)6X]PF6 (X = Cl, Br, I) and [{Mo(CNCMe3)4(NN)}2(μ-CN)](PF6)3 (NN = bpy, Me2bpy, phen) have been studied. The 95Mo chemical shift range for this group of complexes is about 1100 ppm. An increase in the size of the R group attached to the isocyanide ligand generally tends to shield the 95Mo nucleus. Replacement of the isocyanide ligand with a phosphorus ligand also increases the shielding, whereas the replacement of isocyanide with a heterocyclic nitrogen donor leads to deshielding by 800–900 ppm. This group of complexes shows a normal halogen dependence, i.e. replacement of Cl? by Br? and I? increases the shielding of the 95Mo nucleus. The cyano-bridged cations [{Mo(CNCMe3)4(NN)}2(μ-CN)]3+ (NN = bpy, Me2bpy, or phen) show two 95Mo NMR signals, one for the molybdenum coordinated to the carbon of the bridging CN and one for the N-coordinated molybdenum. Comparison of the chemical shifts and linewidths of the cyano-bridged species with those of the corresponding mononuclear molybdenum(II) complexes [Mo(CNCMe3)5(NN)](PF6)2 leads to the assignment of the more deshielded signal to the N-coordinated molybdenum. The 14N and 31P NMR spectra for these complexes have also been measured, as have the 13C NMR spectra of the pairs of complexes [Mo(CNCMe3)5(NN)](PF6)2 and [{Mo(CNCMe3)4(NN)}2(μ-CN)](PF6)3 (NN = bpy or phen). The 183W NMR spectra for [W(CNR)5(bpy)](PF6)2 (R = CMe3 and CH2Ph), show that the δ(183W)/δ(95Mo) chemical shift ratios for isocyanide complexes are different from the ratio found for M0 and MVI.  相似文献   

2.
A copper(II) and two nickel(II) dinuclear oxalato‐bridged compounds of formulae [{Cu(bpdto)}2(μ‐ox)](ClO4)2 ( 1 ), [{Ni(bpdto)]2(μ‐ox)](ClO4)2( 2 ), and [{Ni(bpdto)}2(μ‐ox)](NO3)2·2H2O ( 3 ), where bpdto = 1, 8‐bis(2‐pyridyl)‐3, 6‐dithiaoctane and ox = oxalate = C2O42— anion, have been synthesized and characterized. The crystal structure of 3 was determined by single‐crystal X‐ray analysis. It is a dinuclear complex with i symmetry in which the oxalate ligand is coordinated in bis(didentate) fashion to the inversion centre‐related nickel atoms. The distorted octahedral environment of each nickel atom is completed by two sulphur atoms in the equatorial plane and by two pyridyl nitrogen atoms in axial positions. Magnetic susceptibility measurements over the range 5 — 299K, show antiferromagnetic interactions that are weak in 1 (J = —12.8 cm—1) and strong in 2 and 3 (J = —37.8 and —40.9 cm—1, respectively), which in the case of 3 is in keeping with the observed structural parameters.  相似文献   

3.
The reaction of [CpRu(CH3CN)3]PF6 with the bidentate ligands L-L=1,2-bis(diphenylphosphino)ethane, dppe, and (1-diphenylarsino-2-diphenylphosphino)ethane, dpadppe, affords mononuclear or dinuclear complexes of formula [CpRu(η2-L-L)(CH3CN)]PF6, [{CpRu(CH3CN)2}2(μ-η1:1-L-L)](PF6)2 and [{CpRu(CH3CN)}2(μ-η1:1-L-L)2](PF6)2 (L-L=dppe, dpadppe). All of the compounds are characterized by microanalysis and NMR [1H and 31P{1H}] spectroscopy. The crystal structure of [{CpRu(CH3CN)2}2(μ-η1:1-dppe)](PF6)2 has been determined by X-ray diffraction analysis. The complex exhibits a dppe ligand bridging two CpRu(CH3CN)2 fragments.  相似文献   

4.
Three mixed-valence copper complexes [{Cu(phen)2}2(μ-L)](PF6)2 (where phen = 1,10-phenanthroline, L = 1,4-dicyanamidobenzene (dicyd)), 1,4-dicyanamido-2,5-dimethylbenzene (Me2dicyd) and 1,4-dicyanamido-2,5-dichlorobenzene (Cl2dicyd), and one dinuclear Cu(II) complex [{Cu(phen)2}2(μ-apc)](PF6)3 (where apc = monoanion of 4-azo(phenylcyanamido)benzene) have been prepared and characterized by elemental analysis, IR and electronic absorption spectroscopies and cyclic voltammetry. [{Cu(phen)2}2(μ-apc)](PF6)3 · 2CH3COCH3 crystallized in the triclinic system and both five-coordinate Cu(II) ions in the dinuclear unit are linked through a bridging 4-azo(phenylcyanamido)benzene (apc) ligand. The cyanamide group (NCN) of the bridging ligand is coordinated to Cu(II) ions through the cyano-nitrogen and amido-nitrogen. The bond length between Cu(1) and cyano-nitrogen is slightly larger than that formed by Cu(2) and amido-nitrogen. The angular structural index parameters, τ, for Cu(1) and Cu(2) are 0.9 and 0.5, respectively. The copper(II) atoms display a different geometry with a N5 chromophore group. The intra Cu?Cu separation is 5.156(1) Å. All of the dicyd dinuclear copper complexes show radical anion absorption.  相似文献   

5.
The cations in the solid-state structures of meso-(ΛΔ)-[{Ru(bpy)2}2(μ-bpm)](PF6)4, meso-(ΛΔ)-[{Ru(Me2bpy)2}2(μ-bpm)](tos)4 · 2CH3OH · 4H2O and meso-(ΛΔ)-[{Ru(Me4bpy)2}2(μ-bpm)](tos)4 · 26H2O (bpm = 2,2′-bipyrimidine; bpy = 2,2′-bipyridine; Me2bpy = 4,4′-dimethyl-2,2′-bipyridine; Me4bpy = 4,4′,5,5′-tetramethyl-2,2′-bipyridine; tos = toluene-4-sulfonate anion) exhibit similar features including comparable bond lengths and angles, and metal–metal separations of 5.56–5.59 Å. The counter-ions present in the structures reside in the clefts above and below the plane of the bridging ligand, but show considerable variation in location compared with their known occupancy in solution.  相似文献   

6.
Treatment of [{Me2C6H(CH2PtBu2)2}Rh(η1‐N2)] ( 1a ) with molecular oxygen (O2) resulted in almost quantitative formation of the dioxygen adduct [{Me2C6H(CH2PtBu2)2}Rh(η2‐O2)] ( 2a ). An X‐ray diffraction study of 2a revealed the shortest O? O bond reported for Rh? O2 complexes, indicating the formation of a RhI? O2 adduct, rather than a cyclic RhIII η2‐peroxo complex. The coordination of the O2 ligand in 2a was shown to be reversible. Treatment of 2a with CO gas yielded almost quantitatively the corresponding carbonyl complex [{Me2C6H(CH2PtBu2)2}Rh(CO)] ( 3a ). Surprisingly, treatment of the structurally very similar pincer complex [{C6H3(CH2PiPr2)2)}Rh(η1‐N2)] ( 1b ) with O2 led to partial decomposition, with no dioxygen adduct being observed.  相似文献   

7.
The ability of the oxonitride [{Ti(η5-C5Me5)(μ-O)}33-N)] (1) to act as an organometallic ligand has been studied from both theoretical and experimental points of view. DFT calculations have allowed understanding the electronic structure of 1, and rationalizing its chemical behavior by comparison with the electronic structures of isoelectronic species [{Ti(η5-C5Me5)(μ-O)}33-CH)] and [{Ti(η5-C5Me5)(μ-NH)}33-N)]. Reactions of 1 with different inorganic molecules such as [Mo(CO)3(1,3,5-Me3C6H3)] or AlEt3 have confirmed the possibility of 1 to act as a tridentate or monodentate ligand to give the [{(CO)3Mo}(μ3-O)3{Ti35-C5Me5)33-N)}] (2) and [{Et3Al}(μ3-O){(μ-O)2Ti35-C5Me5)33-N)}] (3) complexes, respectively. Surprisingly, reactions of 1 with [M(CO)6] (M = Cr, Mo, W) complexes led to activate the μ3-N unit in 1 to afford the new compounds [Ti35-C5Me5)3(μ-O)4{(NC)M(CO)5}]2 [M = Cr (4), Mo (5), W (6)]. Molecular structures of complexes 2-6 have been established by single crystal X-ray analysis.  相似文献   

8.
The aqua complexes (SM,RC)-[(η5-C5Me5)M{(R)-Prophos}(H2O)](SbF6)2 (M = Rh, Ir; (R)-Prophos = 1,2-bisdiphenylphosphino propane) catalyze the 1,3-dipolar cycloaddition reaction (DCR) of nitrones with α,β-unsaturated nitriles with low-to-moderate enantioselectivity. The involved catalysts [(η5-C5Me5)M{(R)-Prophos}(α,β-unsaturated nitrile)](SbF6)2, isolated as mixtures of the (SM,RC)- and (RM,RC)-diastereomers, have been fully characterized, and the molecular structure of the complexes (SRh,RC)-[(η5-C5Me5)Rh{(R)-Prophos}(cis-2-pentenenitrile)](SbF6)2 and (SIr,RC)-[(η5-C5Me5)Ir{(R)-Prophos}(acrylonitrile)](SbF6)2 has been determined by X-ray diffraction. The (R)-at-metal epimers isomerize to the (S)-at-metal counterparts. Diastereopure (SM,RC)-[(η5-C5Me5)M{(R)-Prophos}(α,β-unsaturated nitrile)](SbF6)2 complexes catalyze the above-mentioned DCR in a stoichiometric manner with up to 97% ee. The results make clear the influence of the metal configuration on the catalytic stereochemical outcome. The catalysts can be recycled without significant loss of either activity or selectivity.  相似文献   

9.
Titanium(III) complexes containing unprecedented (NH2BH2NHBH3)2− and {N(BH3)3}3− ligands have been isolated, and their structures elucidated by a combination of experimental and theoretical methods. The treatment of the trimethyl derivative [TiCp*Me3] (Cp*=η5-C5Me5) with NH3BH3 (3 equiv) at room temperature gives the paramagnetic dinuclear complex [{TiCp*(NH2BH3)}2(μ-NH2BH2NHBH3)], which at 80 °C leads to the trinuclear hydride derivative [{TiCp*(μ-H)}33-N(BH3)3}]. The bonding modes of the anionic BN fragments in those complexes, as well as the dimethylaminoborane group trapped on the analogous trinuclear [{TiCp*(μ-H)}33-H)(μ3-NMe2BH2)], have been studied by X-ray crystallography and density functional theory (DFT) calculations.  相似文献   

10.
The meso-pyridyl substituted dipyrromethane ligands 5-(4-pyridyl)dipyrromethane (4-dpmane) and 5-(3-pyridyl)dipyrromethane (3-dpmane) have been employed in the synthesis of a series of complexes with the general formulations [(η6-arene)RuCl2(L)] (η6-arene = C6H6, C10H14) and [(η5-C5Me5)MCl2(L)] (M = Rh, Ir). The reaction products have been characterized by microanalyses and spectral studies and molecular structures of the complexes [(η6-C10H14)RuCl2(4-dpmane)] and [(η5-C5Me5)IrCl2(3-dpmane)] have been determined crystallographically. For comparative studies, geometrical optimization have been performed on the complex [(η5-C5Me5)IrCl2(4-dpmane)] using exchange correlation functional B3LYP. Optimized bond length and angles are in good agreement with the structural data of the complex [(η5-C5Me5)IrCl2(3-dpmane)]. The complexes [(η6-C10H14)RuCl2(3-dpmane)], [(η5-C5Me5)RhCl2(3-dpmane)] and [(η5-C5Me5)IrCl2(3-dpmane)] have been employed as a transfer hydrogenation catalyst in the reduction of aldehydes. It was observed that the rhodium and iridium complexes mentioned above are more effective in this regard in comparison to the ruthenium complex.  相似文献   

11.
Substituted-benzoate complexes of nickel(II) of the types bidentate [Ni(mcN3)(Bz)](PF6) and monodentate [Ni(mcN3)(Bz)(H2O)](PF6) have been prepared by acid-base reaction between the hydroxo complexes [Ni(mcN3)(μ-OH)]2(PF6)2 (mcN3 = 2,4,4-trimethyl-1,5,9-triazacyclododec-1-ene (Me3-mcN3) or its 9-methyl derivative (Me4-mcN3)) and the corresponding benzoic acid. The paramagnetic nickel(II) complexes have been characterized in solution by NMR spectroscopy. The influence of the substituents on the hyperfine shift patterns for substituted-benzoate complexes of nickel(II) has been studied. The substituent effects upon the coordination mode of substituted benzoates have been established by X-ray diffraction.  相似文献   

12.
Complexes [{2,6-(Me2NCH2)2C6H3} (p-tolylNYNR)PtHgBrCl] (Y  CH, N; R  Me, Et, i-Pr) have been prepared by the reaction of [{2,6-(Me2NCH2)2C6H3}-PtBr] with [Hg(p-tolylNYNR)Cl]. Similar complexes were obtained, although in lower yields, from exchange reactions of [{2,6-(Me2NCH2)2C6H3} (RCO2)-PtHg(O2CR)Br] with p-tolylNNN(H)-p-tolyl and p-tolylNC(H)N(H)Et.The proposed structure for these heterodinuclear compounds involves a Pt-to-Hg donor bond which is bridged by a triazenido (Y  N) or a formamidino (Y  CH) group, the five-membered ring thus formed acting as a stabilizing factor. The absence of a subsequent electron transfer reaction is ascribed to the constraints of the terdentate 2,6-(Me2NCH2)2C6H3 ligand, which fixes the N-donor atoms in mutual trans-positions.The use of p-tolylNYNR, where R is an alkyl group, results in the formation of two isomers of [{2,6-(Me2NCH2)2C6H3} (p-tolylNYNR)PtHgBrCl] with p-tolyl-N and alkyl-N sites bonded either to Pt or Hg. The relative abundance of these isomers varies systematically with the nature of the group R. It is suggested that the ratio is determined during the formation of the complexes and that both steric and electronic factors are important.  相似文献   

13.
The organolithium reagent [{HC(Ph2PNC6H2Me3-2,4,6)2}Li(OEt2)] was easily obtained by deprotonation of H2C(Ph2PNC6H2Me3-2,4,6)2 with nBuLi in diethyl ether solution. The crystal structure of [{HC(Ph2PNC6H2Me3-2,4,6)2}Li(OEt2)] has been determined and shown to consist a monomeric chelate structure that contains a distorted, trigonal planar lithium centre. The ligand precursor has also been deprotonated with both Me3Al and Me2AlCl to yield the tetrahedral organoaluminium complexes, [{HC(Ph2PNC6H2Me3-2,4,6)2}AlMe2] and [{HC(Ph2PNC6H2Me3-2,4,6)2}Al(Cl)Me]. Reaction of [{HC(Ph2PNC6H2Me3-2,4,6)2}Li(OEt2)] with either AlX3 (X=Cl, Br, I) or GaCl3 yielded a series of dihalo derivatives [{HC(Ph2PNC6H2Me3-2,4,6)2}MX2] all of which have been shown to exist as similar monomeric species containing four-coordinate group 13 centres.  相似文献   

14.
Bismuth compounds are gaining importance as potential alternatives to transition-metal complexes and electron deficient lighter p-block compounds in homogeneous catalysis. Computational analysis on the two-coordinate [(Me2NC6H4)Bi]2+ possessing three electrophilic sites is experimentally evidenced by the isolation of [{Me2NC6H4}Bi{OP(NMe2)3}3][B(3,5-C6H3Cl2)4]2. These observations led us to generate dicationic organobismuth catalyst, [(Me2NC6H4)Bi(L)3]2+ (L=aldehyde/ketone), evidenced by NMR spectroscopy in solution and by single-crystal X-ray diffraction in the solid state. It efficiently catalyzes hydrosilylation of aldehydes and ketones resulting in silyl ethers as the only products in high yields. Our investigations support a carbonyl activation mechanism at the bismuth center followed by Si−H addition.  相似文献   

15.
Summary Four new trinuclear copper(II) complexes, [Cu(phen)-(NBzIm)] (ClO4) (1), [Cu(bpy)(NBzIm)](ClO4) (2), [Cu-(Me2-bpy)(NBzIm)](Ac)·1/2H2O (3) and [Cu(Me2-bpy)-(Im)](ClO4)·1/2H2O (4) (phen = 1, 10-phenanthroline, bpy = 2,2-bipyridine, NBzIm = 6-nitrobenzimidazolate ion, Im=imidazolate ion) have been prepared and characterized by variable temperature magnetic susceptibility measurements. A weak antiferromagnetic spin exchange interaction operates between copper(II) ions, exchange integrals evaluated as J =-23.82 cm-1 for (1); and J=-21.91 cm-1 for (2).  相似文献   

16.
Homoleptic tetramethylaluminate complexes [Ln(AlMe4)3] (Ln=La, Nd, Y) reacted with HCpNMe2 (CpNMe2=1‐[2‐(N,N‐dimethylamino)‐ethyl]‐2,3,4,5‐tetramethyl‐cyclopentadienyl) in pentane at ?35 °C to yield half‐sandwich rare‐earth‐metal complexes, [{C5Me4CH2CH2NMe2(AlMe3)}Ln(AlMe4)2]. Removal of the N‐donor‐coordinated trimethylaluminum group through donor displacement by using an equimolar amount of Et2O at ambient temperature only generated the methylene‐bridged complexes [{C5Me4CH2CH2NMe(μ‐CH2)AlMe3}Ln(AlMe4)] with the larger rare‐earth‐metal ions lanthanum and neodymium. X‐ray diffraction analysis revealed the formation of isostructural complexes and the C? H bond activation of one aminomethyl group. The formation of Ln(μ‐CH2)Al moieties was further corroborated by 13C and 1H‐13C HSQC NMR spectroscopy. In the case of the largest metal center, lanthanum, this C? H bond activation could be suppressed at ?35 °C, thereby leading to the isolation of [(CpNMe2)La(AlMe4)2], which contains an intramolecularly coordinated amino group. The protonolysis reaction of [Ln(AlMe4)3] (Ln=La, Nd) with the anilinyl‐substituted cyclopentadiene HCpAMe2 (CpAMe2=1‐[1‐(N,N‐dimethylanilinyl)]‐2,3,4,5‐tetramethylcyclopentadienyl) at ?35 °C generated the half‐sandwich complexes [(CpAMe2)Ln(AlMe4)2]. Heating these complexes at 75 °C resulted in the C? H bond activation of one of the anilinium methyl groups and the formation of [{C5Me4C6H4NMe(μ‐CH2)AlMe3}Ln(AlMe4)] through the elimination of methane. In contrast, the smaller yttrium metal center already gave the aminomethyl‐activated complex at ?35 °C, which is isostructural to those of lanthanum and neodymium. The performance of complexes [{C5Me4CH2CH2NMe(μ‐CH2)AlMe3}‐ Ln(AlMe4)], [(CpAMe2)Ln(AlMe4)2], and [{C5Me4C6H4NMe(μ‐CH2)AlMe3}Ln(AlMe4)] in the polymerization of isoprene was investigated upon activation with [Ph3C][B(C6F5)4], [PhNMe2H][B(C6F5)4], and B(C6F5)3. The highest stereoselectivities were observed with the lanthanum‐based pre‐catalysts, thereby producing polyisoprene with trans‐1,4 contents of up to 95.6 %. Narrow molecular‐weight distributions (Mw/Mn<1.1) and complete consumption of the monomer suggested a living‐polymerization mechanism.  相似文献   

17.
New series of mono and binuclear arene ruthenium complexes [{(η6-arene)RuCl(L)}]+ and [{(η6-arene)RuCl}2(μ-L)2]2+ (arene=benzene, p-cymene or hexamethylbenzene), {L=pyridine-2-carbaldehyde azine (paa), p-phenylene-bis(picoline)-aldimine (pbp) and p-bi-phenylene-bis(picoline)-aldimine (bbp)} are reported. The complexes have been fully characterized and molecular structure of the representative mononuclear complex [(η6-C6Me6)RuCl(paa)]BF4 (1), binuclear complexes [{(η6-C10H14)RuCl}2(μ-paa)](BF4)2 (3) and [{(η6-C10H14)RuCl}2(μ-pbp)](BF4)2 (6) have been determined by single crystal X-ray diffraction analyses. Single crystal X-ray structure determination revealed that in the binuclear complexes the [(η6-C10H14)RuCl]+ units are trans disposed. Further, the crystal packing in the complexes 1, 3 and 6 is stabilized by C-H?X type (X=Cl, F) inter, intramolecular hydrogen bonding and π-π stacking (3). To explore the ambiguous nature of the bonding between pyridine-2-carbaldehyde azine (paa) with ruthenium containing units [(η6-arene)RuCl]+, DFT/B3LYP calculations have been performed on the complexes [(η6-arene)RuCl(paa)]+ (arene=C6H6, I; C6Me6, II; C10H14, III).  相似文献   

18.
The ligand exchange reaction of IMe-(CH2)2-PPh2 (IMe = 1-methyimidazol-2-ylidene) and the hexacarbonyl complex [{Fe2{μ-S(CH2)3S}(CO)6] (1) resulted in the formation of the chelated complex [{Fe2{μ-S(CH2)3S}(CO)4(IMe-(CH2)2-PPh2)] (2). The molecular structure of 2 was confirmed by spectroscopic and X-ray analyses. This complex catalyzes proton reduction. Low temperature NMR studies on the protonation of 2 revealed the formation of a terminal hydride intermediate.  相似文献   

19.
Heterobimetallic cationic sandwich complexes [M(μ-Cp)M′Cp]+ of group 13 (M=Ga, In) and group 14 (M′=Ge, Sn) elements have been prepared as [WCA] salts (WCA=Al(ORF)4; ORF=OC(CF3)3). Their molecular structures include free apical gallium or indium atoms. The sandwich complexes were formed in the reactions of [M(HMB)]+[WCA] (HMB=C6Me6) with the free metallocenes [M′Cp2]. Their structures are related to known stannocene and stannocenium salts; the unprecedented germanium analogues, namely the free germanocenium cation [GeCp]+ and the corresponding triple-decker complex cation [CpGe(μ-Cp)GeCp]+, are described herein. By variation of the reaction conditions, these sandwich complexes can be transformed into the group 13/14 mixed cationic coordination polymer [{In(HMB)(μ-SnCp2)}n][WCA]n. This polymeric chain motif was also successfully replicated by the synthesis of complexes [{Ga/In(HMB)(μ-FeCp2)}n][WCA]n containing FeCp2 as a bridging ligand.  相似文献   

20.
The reaction of the base‐free terminal thorium imido complex [{η5‐1,2,4‐(Me3C)3C5H2}2Th?N(p‐tolyl)] ( 1 ) with p‐azidotoluene yielded irreversibly the tetraazametallacyclopentene [{η5‐1,2,4‐(Me3C)3C5H2}2Th{N(p‐tolyl)N?N? N(p‐tolyl)}] ( 2 ), whereas the bridging imido complex [{[η5‐1,2,4‐(Me3C)3C5H2]Th(N3)2}2{μ‐N(p‐tolyl)}2][(n‐C4H9)4N]2 ( 3 ) was isolated from the reaction of 1 with [(n‐C4H9)4N]N3. Unexpectedly, upon the treatment of 1 with 9‐diazofluorene, the NN bond was cleaved, an N atom was transferred, and the η2‐diazenido iminato complex [{η5‐1,2,4‐(Me3C)3C5H2}2Th{η2‐[N?N(p‐tolyl)]}{N?(9‐C13H8)}] ( 4 ) was formed. In contrast, the reaction of 1 with Me3SiCHN2 gave the nitrilimido complex [{η5‐1,2,4‐(Me3C)3C5H2}2Th{NH(p‐tolyl)}{N2CSiMe3}] ( 5 ), which slowly converted into [{η5‐1,2,4‐(Me3C)3C5H2}{η5:κ‐N‐1,2‐(Me3C)2‐4‐CMe2(CH2NN?CHSiMe3)C5H2}Th{NH(p‐tolyl)}] ( 6 ) by intramolecular C? H bond activation. The experimental results are complemented by density functional theory (DFT) studies.  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