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.
Singlet cyclo‐1,3‐dipnicta‐2,4‐diazane‐1,3‐diyls of the type [E(μ‐NTer)2E] ( 2 , E=P, As, Ter=2,6‐dimesitylphenyl) can undergo a one‐electron‐oxidation utilizing silver salts of weakly coordinating anions such as [AgLn][B(C6F5)4] (L=donor solvents) to afford the novel cyclic radical cations, [E(μ‐NTer)2E]+. ( 3 +.). When smaller and more basic anions were employed in the reaction, the anions were found to form covalent bonds to the radical centers yielding dipnictadiazanes, [FP(μ‐NTer)2PF] ( 5 ) and [(CF3CO2)P(μ‐NTer)2P(CF3CO2)] ( 6 ). A two‐electron oxidation process, resulting in the formation of dications of the type [E(μ‐NTer)2E]2+, could not be observed. Computational and EPR data revealed that the spin density is almost completely localized at the two heavier pnictogen centers E of the former 1,3‐dipnictadiazane‐1,3‐diyls. The bonding situation in the radical cations features a rare example of a transannular one‐electron π bond without having a σ bond.  相似文献   

2.
A common bridging ligand, 3,3′,5,5′‐tetrakis(N‐methylbenzimidazol‐2‐yl)biphenyl, and four terpyridine terminal ligands with various substituents (amine, tolyl, nitro, and ester groups) have been used to synthesize ten cyclometalated diruthenium complexes 1 2+– 10 2+. Among them, compounds 1 2+– 6 2+ are redox nonsymmetric, and others are symmetric. These complexes show two RuIII/II processes and an intervalence charge transfer (IVCT) transition in the one‐electron oxidized state. The potential separation (ΔE) of 1 2+– 10 2+ has been correlated to the energy difference ΔG0, the energy of the IVCT band Eop, and the ground‐state delocalization coefficient α2. Time‐dependent (TD)DFT calculations suggest that the absorptions in the visible region of 1 2+– 6 2+ are mainly associated with the metal‐to‐ligand charge‐transfer transitions from both ruthenium ions and to both terminal ligands and the bridging ligand. However, the energies of these transitions vary significantly. DFT calculations have been performed on 1 2+– 6 2+ and 1 3+– 6 3+ to give information on the electronic structures and spin populations of the mixed‐valent compounds. The TDDFT‐predicted IVCT excitations reproduce well the experimental trends in transition energies. In addition, three monoruthenium complexes have been synthesized for a comparison study.  相似文献   

3.
Radical cations of a heptathienoacene α,β‐substituted with four n‐decyl side groups (D4T7 . +) form exceptionally stable π‐dimer dications already at ambient temperature (Chem. Comm. 2011 , 47, 12622). This extraordinary π‐dimerization process is investigated here with a focus on the ultimate [D4T7 . +]2 π‐dimer dication and yet‐unreported transitory species formed during and after the oxidation. To this end, we use a joint experimental and theoretical approach that combines cyclic voltammetry, in situ spectrochemistry and spectroelectrochemistry, EPR spectroscopy, and DFT calculations. The impact of temperature, thienoacene concentration, and the nature and concentration of counteranions on the π‐dimerization process is also investigated in detail. Two different transitory species were detected in the course of the one‐electron oxidation: 1) a different transient conformation of the ultimate [D4T7 . +]2 π‐dimer dications, the stability of which is strongly affected by the applied experimental conditions, and 2) intermediate [D4T7]2 . + π‐dimer radical cations formed prior to the fully oxidized [D4T7]2 . + π‐dimer dications. Thus, this comprehensive work demonstrates the formation of peculiar supramolecular species of heptathienoacene radical cations, the stability, nature, and structure of which have been successfully analyzed. We therefore believe that this study leads to a deeper fundamental understanding of the mechanism of dimer formation between conjugated aromatic systems.  相似文献   

4.
Reactions of the dimeric cobalt complex [(L?Co)2] ( 1 , L=[(2,6‐iPr2C6H3)NC(Me)]2) with polyarenes afforded a series of mononuclear and dinuclear complexes: [LCo(η4‐anthracene)] ( 2 ), [LCo(μ‐η44‐naphthalene)CoL] ( 3 ), and [LCo(μ‐η44‐phenanthrene)CoL] ( 4 ). The pyrene complexes [{Na2(Et2O)2}{LCo(μ‐η33‐pyrene)CoL}] ( 5 ) and [{Na2(Et2O)3}{LCo(η3‐pyrene)}] ( 6 ) were obtained by treating precursor 1 with pyrene followed by reduction with Na metal. These complexes contain three potential redox active centers: the cobalt metal and both α‐diimine and polyarene ligands. Through a combination of X‐ray crystallography, EPR spectroscopy, magnetic susceptibility measurement, and DFT computations, the electronic configurations of these complexes were studied. It was determined that complexes 2 – 4 have a high‐spin CoI center coupled with a radical α‐diimine ligand and a neutral polyarene ligand. Whereas, the ligand L in complexes 5 and 6 has been further reduced to the dianion, the cobalt remains in a formal (I) oxidation state, and the pyrene molecule is either neutral or monoanionic.  相似文献   

5.
Methionine (Met) is an essential sulfur‐containing amino acid, sensitive to oxidation. The oxidation of Met can occur by numerous pathways, including enzymatic modifications and oxidative stress, being able to cause relevant alterations in protein functionality. Under UV radiation, Met may be oxidized by direct absorption (below 250 nm) or by photosensitized reactions. Herein, kinetics of the reaction and identification of products during photosensitized oxidation were analyzed to elucidate the mechanism for the degradation of Met under UV‐A irradiation using pterins, pterin (Ptr) and 6‐methylpterin (Mep), as sensitizers. The process begins with an electron transfer from Met to the triplet‐excited state of the photosensitizer (Ptr or Mep), to yield the corresponding pair of radicals, Met radical cation (Met?+) and the radical anion of the sensitizer (Sens??). In air‐equilibrated solutions, Met?+ incorporates one or two atoms of oxygen to yield methionine sulfoxide (MetO) and methionine sulfone (MetO2), whereas Sens?? reacts with O2 to recover the photosensitizer and generate superoxide anion (O2??). In anaerobic conditions, further free‐radical reactions lead to the formation of the corresponding dihydropterin derivatives (H2Ptr or H2Mep).  相似文献   

6.
The bis(vinyl ruthenium)‐modified squaraine dye 1 was synthesized by treatment of [RuHCl(CO)(PiPr3)2] with bis(ethynyl)‐substituted squaraine 8 . Spectroscopic and electrochemical measurements on 1 and its organic precursors 6 – 8 were performed to study the effect of the vinyl ruthenium “substituents,” particularly with respect to (poly)electrochromism. Attachment of the vinyl ruthenium moieties endows metal–organic squaraine 1 with two additional oxidation waves and lowers the first two oxidation potentials by approximately 300 mV with respect to its organic precursors. Squaraines 6 , 7 , 8 , and 1 strongly absorb at 648, 663, 656, or 709 nm. Although organic dyes 6 , 7 , and 8 fluoresce, no room‐temperature emission is observed for 1 . The radical cations and anions of 6 , 7 , 8 , and 1 as well as the doubly oxidized dications have been studied by IR and UV/Vis/NIR spectroelectrochemistry, and the ?/0/+/2+ redox sequences were found to be reversible in each case. Our results indicate that the 12?/?/0/+/2+ redox system constitutes a polyelectrochromic switch in which absorption in the visible or the near‐infrared range is reversibly turned off or shifted deep into the NIR. They also show that radical cation 1.+ is an intrinsically delocalized system with only little contribution from the outer vinyl ruthenium tags to the oxidation process. Dication 12+ constitutes a class‐II mixed‐valent system with two electronically different vinyl ruthenium moieties and has an open‐shell singlet electronic ground‐state structure. ESR and NMR spectra of chemically prepared 1.+ and 12+ corroborate these results. It has also emerged that reduction involves an orbital that is strongly delocalized across the entire squaraine π system and strongly affects the peripheral vinyl ruthenium sites.  相似文献   

7.
A charge-transfer (CT) complex of NOBF4 and hexamethoxybenzene (HMB), which gives out HMB?+ as a “fluorescent radical cation probe,” upon one-electron oxidation, has been designed to explore the excited state dynamics of contact radical ion pairs by laser-induced fluorescence and femtosecond transient absorption spectroscopic techniques. The acetonitrile solution of the CT complex showed weak fluorescence with a similar spectrum to that observed for free excited HMB radical cation (HMB?+*), suggesting the formation of HMB?+* upon the one-photonic excitation of the CT complex. The laser-power dependence of the fluorescence intensity supported the one-photonic excitation event. We have also observed a short-lived transient species but no long-lived species by femtosecond laser flash photolysis of the CT complex. The lifetime (6.5 ps) was in good accordance with its fluorescence quantum yield (2.5 × 10?5) and was able to assign the transient species to the fluorescent state, an excited radical ion pair [HMB ?+*/NO?]. All the events were completed within the inner sphere and the short lifetime of the transient species could be attributed to rapid back-electron transfer. It is concluded that the excited radical cation character in the excited state of the CT complex originates from the radical ion character in the CT complex in the ground state and that a relatively long lifetime of HMB?+* facilitates its observation even in the contact ion pair.  相似文献   

8.
Photodissociation of pyrene (Py) dimer radical cation (Py 2 ?+ ) giving pyrene radical cation (Py?+) and Py and subsequent regeneration of Py 2 ?+ by association of Py?+ and Py were directly observed during the pulse radiolysis–laser flash photolysis combined method at room temperature. When Py 2 ?+ was excited at the local excitation band with the 532-nm laser flash, the rapid growth and decay of monomeric Py?+ were observed at 460 nm. The dissociation of Py 2 ?+ proceeded via a one-photon process to give the ground-state Py?+(D0) and Py in the quantum yield (Φdiss) of (2.9 ± 0.9) × 10?3. It was shown that Py?+ decayed with a time constant of several tens of nanoseconds, indicating that the association of Py?+ with Py regenerating Py 2 ?+ proceeds at a diffusion-controlled rate. The photodissociation proceeded from the lowest excited state of Py 2 ?+ , even when Py 2 ?+ was excited to the higher excited state. The difference between the Φdiss value of Py 2 ?+ and that previously reported for naphthalene dimer radical cation (Np 2 ?+ ) is discussed.  相似文献   

9.
The divinyldiarsene radical cations [{(NHC)C(Ph)}As]2(GaCl4) (NHC=IPr: C{(NDipp)CH}2 3 ; SIPr: C{(NDipp)CH2}2 4 ; Dipp=2,6‐iPr2C6H3) and dications [{(NHC)C(Ph)}As]2(GaCl4)2 (NHC=IPr 5 ; SIPr 6 ) are readily accessible as crystalline solids on sequential one‐electron oxidation of the corresponding divinyldiarsenes [{(NHC)C(Ph)}As]2 (NHC=IPr 1 ; SIPr 2 ) with GaCl3. Compounds 3 – 6 have been characterized by X‐ray diffraction, cyclic voltammetry, EPR/NMR spectroscopy, and UV/vis absorption spectroscopy as well as DFT calculations. The sequential removal of one electron from the HOMO, that is mainly the As?As π‐bond, of 1 and 2 leads to successive elongation of the As=As bond and contraction of the C?As bonds from 1 / 2 → 3 / 4 → 5 / 6 . The UV/vis spectrum of 3 and 4 each exhibits a strong absorption in the visible region associated with SOMO‐related transitions. The EPR spectrum of 3 and 4 each shows a broadened septet owing to coupling of the unpaired electron with two 75As (I=3/2) nuclei.  相似文献   

10.
The complex series [Ru(pap)(Q)2]n ([ 1 ]n–[ 4 ]n; n=+2, +1, 0, ?1, ?2) contains four redox non‐innocent entities: one ruthenium ion, 2‐phenylazopyridine (pap), and two o‐iminoquinone moieties, Q=3,5‐di‐tert‐butyl‐N‐aryl‐1,2‐benzoquinonemonoimine (aryl=C6H5 ( 1+ ); m‐(Cl)2C6H3 ( 2+ ); m‐(OCH3)2C6H3 ( 3+ ); m‐(tBu)2C6H3 ( 4 +)). A crystal structure determination of the representative compound, [ 1 ]ClO4, established the crystallization of the ctt‐isomeric form, that is, cis and trans with respect to the mutual orientations of O and N donors of two Q ligands, and the coordinating azo N atom trans to the O donor of Q. The sensitive C? O (average: 1.299(3) Å), C? N (average: 1.346(4) Å) and intra‐ring C? C (meta; average: 1.373(4) Å) bond lengths of the coordinated iminoquinone moieties in corroboration with the N?N length (1.292(3) Å) of pap in 1 + establish [RuIII(pap0)(Q.?)2]+ as the most appropriate electronic structural form. The coupling of three spins from one low‐spin ruthenium(III) (t2g5) and two Q.? radicals in 1 +– 4 + gives a ground state with one unpaired electron on Q.?, as evident from g=1.995 radical‐type EPR signals for 1 +– 4 +. Accordingly, the DFT‐calculated Mulliken spin densities of 1 + (1.152 for two Q, Ru: ?0.179, pap: 0.031) confirm Q‐based spin. Complex ions 1 +– 4 + exhibit two near‐IR absorption bands at about λ=2000 and 920 nm in addition to intense multiple transitions covering the visible to UV regions; compounds [ 1 ]ClO4–[ 4 ]ClO4 undergo one oxidation and three separate reduction processes within ±2.0 V versus SCE. The crystal structure of the neutral (one‐electron reduced) state ( 2 ) was determined to show metal‐based reduction and an EPR signal at g=1.996. The electronic transitions of the complexes 1 n– 4 n (n=+2, +1, 0, ?1, ?2) in the UV, visible, and NIR regions, as determined by using spectroelectrochemistry, have been analyzed by TD‐DFT calculations and reveal significant low‐energy absorbance (λmax>1000 nm) for cations, anions, and neutral forms. The experimental studies in combination with DFT calculations suggest the dominant valence configurations of 1 n– 4 n in the accessible redox states to be [RuIII(pap0)(Q.?)(Q0)]2+ ( 1 2+– 4 2+)→[RuIII(pap0)(Q.?)2]+ ( 1 +– 4 +)→[RuII(pap0)(Q.?)2] ( 1 – 4 )→[RuII(pap.?)(Q.?)2]? ( 1 ?– 4 ?)→[RuIII(pap.?)(Q2?)2]2? ( 1 2?– 4 2?).  相似文献   

11.
A trefoil‐like porphyrin trimer linked by triphenylamine (TPA‐TPZn3) was synthesized. A three‐electron oxidation of TPA‐TPZn3 forms a radical trication (TPA‐TPZn33+), in which each porphyrin ring undergoes a one‐electron oxidation. The radical trication TPA‐TPZn33+ spontaneously dimerizes to afford (TPA‐TPZn3)26+ in CH2Cl2. The characteristic charge‐resonance band due to the charge delocalization over the π system of (TPA‐TPZn3)26+ was observed in the NIR region. The initial oxidation potential of TPA‐TPZn3 is negatively shifted relative to that of the corresponding monomer porphyrin, which results from the stabilization of the oxidized state of TPA‐TPZn3 associated with the dimerization. The thermodynamic parameters (i.e., ΔH, ΔS, and ΔG) for the formation of (TPA‐TPZn3)26+ were determined by measuring Vis/NIR spectra at various temperatures. The formation constant of (TPA‐TPZn3)26+ is significantly larger than that of the radical cation dimer of the corresponding monomer porphyrin (e.g., over 2000‐fold at 233 K). The electronic states were investigated using EPR spectroscopic analysis. The greatly enhanced dimerization of TPA‐TPZn33+ results from multiple π‐bond formation between the porphyrin radical cations.  相似文献   

12.
A series of chiral Schiff bases ( L 1 – L 5 ) with different substituents in the salicylidenyl unit were prepared from condensation of 3‐aryl‐5‐ tert ‐butylsalicylaldehyde derivatives and optically active amino alcohols. Bromination of 3‐phenyl‐5‐ tert ‐butylsalicylaldehyde gave an unexpected product 3‐(4‐bromophenyl)‐5‐bromosalicylaldehyde, from which the corresponding Schiff base ligands L 6 and L 7 , derived from (S)‐valinol and (S)‐ tert ‐leucinol, respectively, were prepared. Ligands L 1 – L 7 were applied to the vanadium‐catalyzed asymmetric oxidation of aryl methyl sulfides. Under the optimal conditions, the oxidation of the thioanisole with H2O2 as oxidant in CH2Cl2 catalyzed by VO(acac)2‐ L 1 – L 7 gives good yields (74–83%) with moderate enantioselectivity (58–77% ee). Ligand L 7 , containing a 4‐bromophenyl group on the 3‐position and a Br atom on the 5‐position of the salicylidenyl moiety, displays an 80–90% ee for vanadium‐catalyzed oxidation of methyl 4‐bromophenyl sulfide and methyl 2‐naphthyl sulfide. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

13.
Exciplexes of 9,10‐dicyanoanthracene (DCA) with alkylbenzene donors in cyclohexane show structureless emission spectra, typical of exciplexes with predominantly charge‐transfer (CT) character, when the donor has a relatively low oxidation potential (Eox), e.g. hexamethylbenzene (HMB). With increasing Eox and stronger mixing with a locally excited (LE) state, vibrational structure begins to appear with 1,2,3,5‐tetramethylbenzene and becomes prominent with p‐xylene (p‐Xy). A simple theoretical model reproduces the spectra and the radiative rate constants, and it reveals several surprises: Even in this nonpolar solvent, the fractional CT character of a highly mixed exciplex varies widely in response to fluctuations in the microscopic environment. Environments that favor the LE (or CT) state contribute more to the blue (or red) side of the overall spectrum. It is known that sparsely substituted benzene radical cations, e.g., p‐Xy?+, are stabilized more in acetonitrile than the heavily substituted HMB?+. Remarkably, ion pairing with DCA?– in cyclohexane leads to even larger differences in the stabilization of these radical cations. The spectra of the low‐Eox donors are almost identical except for displacements that approximately equal the differences in Eox, even though the exciplexes have varying degrees of CT character. These similarities result from compensation among several nonobvious, but quantified factors.  相似文献   

14.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  相似文献   

15.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  相似文献   

16.
EPR study showed that the semi‐quinone radical anion of 2,3‐dicyano‐5,6‐dichloro‐1,4‐benzoquinone (DDQ) was formed in a charge transfer process between ground‐state DDQ as acceptor and each one of following ground state donors, i.e., 4‐methyl‐4′‐tridecyl‐2, 2′‐bipyridyl; 4‐methyl‐4′‐nonyl‐2, 2′‐bipyridyl; bis (2,2′‐bipyridyl) (4‐methyl‐4′‐heptadecyl‐2, 2′‐bipyridyl)ruthenium(2+) perchlorate and perylene. EPR study also showed that there are perylene cation radical and pyrene cation radical in the following experimental conditions: (a) in 98% sulfuric add. (b) 10?3 mol/L perylene (or pyrene) was dissolved in trifluoroacetic acid‐nitrobenzene (1: 1 V/V).  相似文献   

17.
More than 80 years after Paneth’s report of dimethyl bismuth, the first monomeric BiII radical that is stable in the solid state has been isolated and characterized. Reduction of the diamidobismuth(III) chloride Bi(NONAr)Cl (NONAr=[O(SiMe2NAr)2]2−; Ar=2,6‐iPr2C6H3) with magnesium affords the BiII radical .Bi(NONAr). X‐ray crystallographic measurements are consistent with a two‐coordinate bismuth in the +2 oxidation state with no short intermolecular contacts, and solid‐state SQUID magnetic measurements indicate a paramagnetic compound with a single unpaired electron. EPR and density functional calculations show a metal‐centered radical with >90 % spin density in a p‐type orbital on bismuth.  相似文献   

18.
Metallamacrocylic tetraruthenium complexes were generated by treatment of 1,4‐divinylphenylene‐bridged diruthenium complexes with functionalized 1,3‐benzene dicarboxylic acids and characterized by HR ESI‐MS and multinuclear NMR spectroscopy. Every divinylphenylene diruthenium subunit is oxidized in two consecutive one‐electron steps with half‐wave potential splittings in the range of 250 to 330 mV. Additional, smaller redox‐splittings between the +/2+ and 0/+ and the 3+/4+ and 2+/3+ redox processes, corresponding to the first and the second oxidations of every divinylphenylene diruthenium entity, are due to electrostatic effects. The lack of electronic coupling through bond or through space is explained by the nodal properties of the relevant molecular orbitals and the lateral side‐by‐side arrangement of the divinylphenylene linkers. The polyelectrochromic behavior of the divinylphenylene diruthenium precursors is retained and even amplified in these metallamacrocyclic structures. EPR studies down to T=4 K indicate that the dications 1‐H2+ and 1‐OBu2+ are paramagnetic. The dications and the tetracation of macrocycle 3‐H display intense (dications) or weak ( 3‐H4+ ) EPR signals. Quantum chemical calculations indicate that the four most stable conformers of the macrocycles are largely devoid of strain. Bond parameters, energies as well as charge and spin density distributions of model macrocycle 5‐HMe were calculated for the different charge and spin states.  相似文献   

19.
The dimesitylnickel(II) complex [(bpy)NiMes2] (Mes = mesityl = 2,4,6‐trimethylphenyl) was prepared and examined spectroscopically and electrochemically. The crystal and molecular structure was determined from single crystal X‐Ray diffraction experiments (monoclinic, P21/n, Z = 4, a = 8.3092(8) Å, b = 18.233(2) Å, c = 15.226(2) Å, β = 98.035(6)°). The nickel atom displays a distorted square planar environment. The axial positions of the square plane are shielded by each one of the methyl groups on the mesityl substituents. The complex shows electrochemical reduction processes that are mainly centered on the bpy ligand as inferred from spectroelectrochemical investigations (EPR and UV/Vis/NIR absorption) of the radical anion or dianion. The observed oxidation is assigned to a NiII/NiIII couple. The title complex exhibits strongly solvatochromic longwavelength electronic absorptions.  相似文献   

20.
The synthesis and comprehensive characterization of the first dicationic tellurium analogues of N‐heterocyclic carbenes (NHCs) has been reported, in both the +2 and +4 oxidation states. For the +2 oxidation state, a base‐stabilized form of TeCl2 is used as the starting material. The dications are isolated by means of halide metathesis and the solid‐state structures confirm the previously calculated diimine bonding arrangement. For TeIV, a diamine is used in a high‐yielding dehydrohalogen coupling reaction from TeCl4. The dicationic NHC analogue is isolated in a base‐stabilized form through halide abstraction and subsequent coordination by pyridine.  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