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 15 毫秒
1.
Electronic structures of binuclear ruthenium complexes [Ru2(terpy)2(tppz)]4+ ( 1A ) and [Ru2Cl2(L)2(tppz)]2+ {L = bpy ( 2A ), phen ( 3A ), and dpphen ( 4A )} were studied by density functional theory calculations. Abbreviations of the ligands (Ls) are bpy = 2,2′‐bipyridine, phen = 1,10‐phenanthroline, dpphen = 4,7‐diphenyl‐1,10‐phenanthroline, terpy = 2,2′:6′,2″‐terpyridine, and tppz = tetrakis(2‐pyridyl)pyrazine. Their mononuclear reference complexes [Ru(terpy)2]2+ ( 1B ) and [RuClL(terpy)]+ {L = bpy ( 2B ), phen ( 3B ), and dpphen ( 4B )} were also examined. Geometries of these mononuclear and binuclear Ru(II) complexes were fully optimized. Their geometric parameters are in good agreement with the experimental data. The binuclear complexes were characterized by electrospray ionization mass spectrometry, UV–Vis spectroscopy, and cyclic voltammograms. Hexafluorophosphate salts of binuclear ruthenium complexes of 3A and 4A were newly prepared. The crystal structure of binuclear complex 1A (PF6)4 was also determined. Orbital interactions were analyzed to characterize the metal‐to‐ligand charge‐transfer (MLCT) states in these complexes. The Cl? ligand works to raise the orbital energy of the metal lone pair, which leads to the low MLCT state. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

2.
The reaction of phenols with the excited state, *[Ru(bpy)3]2+ (E0 = 0.76 V) and *[Ru(H2dcbpy)3]2+, (dcbpy = 4,4′‐dicarboxy‐2,2′‐bipyridine) (E0 = 1.55 V vs. SCE) complexes in CH3CN has been studied by luminescence quenching technique and the quenching is dynamic. The formation of phenoxyl radical as a transient is confirmed by its characteristic absorption at 400 nm. The kq value is highly sensitive to the change of pH of the medium and ΔG0 of the reaction. Based on the treatment of kq data in terms of energetics of the reaction and pH of the medium, proton coupled electron transfer (PCET) mechanism has been proposed for the reaction. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

3.
A series of 1,3‐bis(2′‐hydroxyethyl)imidazolium ionic liquids is reported where 1H NMR chemical shift values and thermal stabilities (Td), as determined by thermogravimetric analysis, are correlated with the hydrogen bonding capability of various anions ([Cl?], [Br?], [CF3CO2?], [NO2?], [MsO?], [NO3?], [TfO?], [BF4?], [NTf2?], and [PF6?]). Use of anions with the strongest hydrogen bonding capability, such as chloride [Cl?], bromide [Br?], and trifluoroacetate [CF3CO2?], led to the furthest observed downfield chemical shift values in DMSO‐d6 and the poorest thermal stabilities ([CF3CO2?] < 200 °C). Thermal stabilities in excess of 350 °C and upfield chemical shift values were observed for ionic liquids, which employed the weakly coordinating triflate [OTf?], tetrafluoroborate [BF4?], or bis(trifluoromethylsulfonyl)imide [NTf2?] anion. Optimized structures of selected ionic liquids, as determined by density functional theory calculations at the B3LYP/6‐31G + (d,p) level, indicated that the anion preferred to be located above the imidazolium ring and in close proximity to the hydroxyl groups. Calculated dissociation energies (ΔE) and a comparison of key bonding distances (C2―H, (C2)H···X, O―H, and (O)H···X) also confirmed this structural preference. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
The Raman and infrared spectra of fac ‐tris(2‐phenylpyridinato‐N,C2′)iridium(III), Ir(ppy)3 and surface‐enhanced resonance Raman spectra of bis(2‐phenyl pyridinato‐) (2,2′bipyridine) iridium (III), [Ir(ppy)2 (bpy)]+ cation were recorded in the wavenumber range 150–1700 cm−1, and complete vibrational analyses of Ir(ppy)3 and [Ir(ppy)2 (bpy)]+ were performed. Most of the vibrational wavenumbers were calculated with density‐functional theory agree with experimental data. On the basis of the results of calculation and comparison of the spectra of both complexes and their analogue [Ru(bpy)3]2+, we assign the vibrational wavenumbers for metal–ligand modes; metal–ligand stretching wavenumbers are 277/307 and 261/236 cm−1 for Ir(ppy)3, and 311/324, 257/270, 199/245 cm−1 for [Ir(ppy)2 bpy]+. Surface‐enhanced Raman scattering spectra of [Ir(ppy)2 bpy]2+ were measured at two wavelengths on the red and blue edges of the low‐energy metal‐to‐ligand charge‐transfer band. According to the enhanced Raman intensities for the vibrational modes of both ligands ppy and bpy, the unresolved charge‐transfer band is deduced to consist of charge‐transfer transitions from the triplet metal to both ligands ppy and bpy. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
Gas‐phase structure, hydrogen bonding, and cation–anion interactions of a series of 1‐(2‐hydroxyethyl)‐3‐methylimidazolium ([HOEMIm]+)‐based ionic liquids (hereafter called hydroxyl ILs) with different anions (X = [NTf2], [PF6], [ClO4], [BF4], [DCA], [NO3], [AC] and [Cl]), as well as 1‐ethyl‐3‐methylimizolium ([EMIm]+)‐based ionic liquids (hereafter called nonhydroxyl ILs), were investigated by density functional theory calculations and experiments. Electrostatic potential surfaces and optimized structures of isolated ions, and ion pairs of all ILs have been obtained through calculations at the Becke, three‐parameter, Lee–Yang–Parr/6‐31 + G(d,p) level and their hydrogen bonding behavior was further studied by the polarity and Kamlet–Taft Parameters, and 1H‐NMR analysis. In [EMIm]+‐based nonhydroxyl ILs, hydrogen bonding preferred to be formed between anions and C2–H on the imidazolium ring, while in [HOEMIm]+‐based hydroxyl ILs, it was replaced by a much stronger one that preferably formed between anions and OH. The O–H···X hydrogen bonding is much more anion‐dependent than the C2–H···X, and it is weakened when the anion is changed from [AC] to [NTf2]. The different interaction between [HOEMIm]+ and variable anion involving O–H···X hydrogen bonding resulted in significant effect on their bulk phase properties such as 1H‐NMR shift, polarity and hydrogen‐bond donor ability (acidity, α). Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

6.
A new derivative of the previously reported 1,2‐bis(benzimidazol‐2‐yl)ethane motif, cation [1H2]2+, was synthesized under microwave irradiation and fully characterized by solution NMR, high‐resolution mass spectrometry, cyclic voltammetry and X‐ray crystallography. This cation presents a linear geometry and incorporates nitro substituents as electrochemical handles. In solution, cation [1H2]2+, is capable of threading the cavity of dibenzo‐24‐crown‐8 ether host (DB24C8) giving rise to a [2]pseudorotaxane complex [1H2?DB24C8]2+, regardless of the counterion, [CF3SO3]? or [CF3COO] ?. The interpenetrated structure of [1H2?DB24C8]2+ was proven by solution NMR and X‐ray crystallography. This host–guest complex is held together by several non‐covalent interactions, such as hydrogen bonding and ion‐dipole. An electrochemical study of [1H2]2+ in the presence of variable amounts of DB24C8 was performed; due to the irreversible redox behavior of cation [1H2]2+, it was not possible to electrochemically control the association/dissociation process with DB24C8. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

7.
Hydrolytic reactions of cyclic bis(3′‐5′)diadenylic acid (c‐di‐AMP) have been followed by Reversed phase high performance liquid chromatography (RP‐HPLC) over a wide pH range at 90 °C. Under neutral and basic conditions (pH ≥ 7), disappearance of the starting material (first‐order in [OH?]) was accompanied by formation of a mixture of adenosine 2′‐monophosphate and 3′‐monophosphate (2′‐AMP and 3′‐AMP). Under very acidic conditions (from H0 = ?0.7 to 0.2), c‐di‐AMP undergoes two parallel reactions (first‐order in [H+]): the starting material is cleaved to 2′‐AMP and 3′‐AMP and depurinated to adenine (i.e., cleavage of the N‐glycosidic bond), the former reaction being slightly faster than the latter one. At pH 1–3, isomerization to cyclic bis(2′‐5′)diadenylic acid competes with the depurination. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

8.
A series of metal‐free compounds, ie, planar triprotonated triazine, triazineH3Cl(PF6)2 ( 1 ), planar triprotonated triazineH3Br(PF6)2 ( 2 ), and nonplanar monoprotonated triazineHPF6 ( 3 ), were prepared. Abbreviations used are triazine = tri‐2‐pyridyltriazine. Ruthenium complexes [RuCl(bpy)(L)](PF6), [RuCl(bpy)(L)](PF6)2, and [Ru(L)2](PF6)2 were also prepared, where bpy is 2,2′‐bipyridine and L's are triazine ( 4 ) and monoprotonated triazine ( 5 ), respectively. Ruthenium complexes [Ru(triazine)2](PF6)2 ( 6 ) were also prepared and crystallized. The X‐ray crystal structures of the 3 compounds 1 , 2 , and 3 and the complex 6 were determined. They were also characterized by electrospray ionization mass spectrometry, UV‐vis spectroscopy, and density functional theory calculations.  相似文献   

9.
The kinetics of oxidation of methylxanthine drug, theophylline (TP), by diperiodatocuprate(III) (DPC) has been investigated in the absence and presence of ruthenium(III) (Ru(III)) as homogeneous catalyst in alkaline medium at a constant ionic strength of 0.21 mol dm?3 spectrophotometrically. The reaction exhibits 1:4 stoichiometry ([TP] : [DPC]) in both the cases. The order of the reaction with respect to [DPC] was unity, while the order with respect to [TP] was less than unity over the concentration range studied in both the cases. The rate was increased with an increase in [OH?] and decreased with an increase in [IO4?]. The order with respect to [Ru(III)] was unity. The ionic strength and dielectic constant of the medium did not affect the rate significantly. The main product 1‐methyl‐(3‐N‐formyl)‐2,4‐purinodione was identified by spot tests, Fourier transform infrared spectroscopy and liquid chromatography–mass spectrometry spectral studies. Based on the experimental results, the possible mechanisms were proposed. The reaction constants involved in the different steps of the mechanisms were evaluated. The catalytic constant (Kc) was also calculated for Ru(III) catalysis at different temperatures. The activation parameters with respect to the catalyst and slow step of the mechanisms were computed, and thermodynamic quantities were determined. Kinetic studies suggest that the active species of DPC and Ru(III) are found to be [Cu(H2IO6)(H2O)2] and [Ru(H2O)5OH]2+, respectively. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

10.
A novel fiber-optic evanescent wave sensor (FOEWS) for O2 detection based on [Ru(bpy)3]2+-doped hybrid fluorinated ORMOSILs (organically modified silicates) has been developed. The sensing element was fabricated by dip-coating the optical fiber with [Ru(bpy)3]2+-doped hybrid fluorinated ORMOSILs composed of n-propyltrimethoxysilane (n-propyl-TriMOS) and 3, 3, 3-trifluoropropyltrimethoxysilane (TFP–TriMOS). Fluorophores of [Ru(bpy)3]2+ were excited by the evanescent wave field produced on the fiber core surface and the emission fluorescence was quenched by O2. Spectroscopic properties have been characterized by FTIR and UV–VIS absorption measurements. By using the presented hybrid fluorinated ORMOSILs, which enhances the coating surface hydrophobicity, the quenching response is increased. The sensitivity of the sensor is 7.5, which is quantified in terms of the ratio I N2/I O2 (I N2 and I O2 represent the fluorescence intensities in pure N2 and pure O2 environments, respectively). The limit of detection (L.O.D.) is 0.01% (3σ) and the response time is about 1 s. Meanwhile, the proposed FOEWS has the advantages of easy fabrication, low cost, fast response and suitable sensitivity for oxygen monitoring using a cheap blue LED as light source and coupling a miniature PMT detector directly to the optical fiber probe.  相似文献   

11.
Ru(II) complexes with weak ligand fields may undergo light-induced ligand dissociation, and the resulted Ru(II) aqua complexes may bind with biomolecules such as DNA, showing potential as photoactivated chemotherapy (PACT) agents. However, Ru(II) complexes with efficient PACT activity are still rare. Some Ru(II) complexes exhibit efficient photoinduced ligand dissociation but poor cytotoxicity. It is speculated that the low nuclear accumulation levels may account for their low PACT efficacy. In order to confirm this hypothesis, the almost noncytotoxic [Ru(7-OCH3-dppz)(4-OCH3-py)4](PF6)2 (Ru1) is loaded on nucleus-targeted C5N2 nanoparticles (NPs). Compared with the free Ru1, Ru1–C5N2 NPs exhibit significantly increased cellular uptake and nuclear accumulation. Therefore, Ru1–C5N2 NPs show efficient PACT activity toward various cancer cell lines (including cisplatin-resistant one) with half maximal inhibitory concentration (IC50) values of 0.18 × 10−6–0.29 × 10−6 m and phototoxicity index (IC50dark/IC50light) values above 137 under both normoxic and hypoxic conditions. Moreover, Ru1–C5N2 NPs also exhibit efficient PACT activity toward cisplatin-resistant 3D multicellular tumor spheroids upon two-photon irradiation (800 nm). The same strategy is also feasible to greatly improve the PACT activity of [Ru(7-OCH3-dppz)(py)4]2+, which itself only has a medium effect. The results may provide new sights for developing efficient Ru(II) PACT agents.  相似文献   

12.
The Ru(III)/Os(VIII)/Pd(II)/Pt(IV)‐catalysed kinetics of oxidation of glycyl–glycine (Gly‐Gly) by sodium N‐chloro‐p‐ toluenesulfonamide (chloramine‐T; CAT) in NaOH medium has been investigated at 308 K. The stoichiometry and oxidation products in each case were found to be the same but their kinetic patterns observed are different. Under comparable experimental conditions, the oxidation‐kinetics and mechanistic behaviour of Gly‐Gly with CAT in NaOH medium is different for each catalyst and obeys the underlying rate laws:
  • Rate = k [CAT]t [Gly‐Gly]0 [Ru(III)][OH?]x
  • Rate = k [CAT]t[Gly‐Gly]x [Os(VIII)]y[OH?]z
  • Rate = k [CAT]t[Gly‐Gly]x [Pd(II)][OH?]y
  • Rate = k [CAT]t[Gly‐Gly]0 [Pt(IV)]x[OH?]y
Here, and x, y, z < 1 in all the cases. The anion of CAT, CH3C6H4SO2NCl?, has been postulated as the common reactive oxidising species in all the cases. Under comparable experimental conditions, the relative ability of these catalysts towards oxidation of Gly‐Gly by CAT are in the order: Os(VIII) > Ru(III) > Pt(IV) > Pd(II). This trend may be attributed to the different d‐electronic configuration of the catalysts. Further, the rates of oxidation of all the four catalysed reactions have been compared with uncatalysed reactions, under identical experimental conditions. It was found that the catalysed reaction rates are 7‐ to 24‐fold faster. Based on the observed experimental results, detailed mechanistic interpretation and the related kinetic modelling have been worked out for each catalyst. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

13.
The three Ru(II) complexes of [Ru(phen)2dppca]2+ (1) [Ru(bpy)2dppca]2+ (2) and [Ru(dmb)2dppca]2+ (3) (where phen = 1,10 phenanthroline, bpy = 2,2-bipyridine, dmb = 2 ,2-dimethyl 2′,2′-bipyridine and polypyridyl ligand containing a single carboxylate functionality dppca ligand (dipyridophenazine-11-carboxylic acid) have been synthesized and characterized. These complexes have been shown to act as promising calf thymus DNA intercalators and a new class of DNA light switches, as evidenced by UV-visible and luminescence titrations with Co2+ and EDTA, steady-state emission quenching by [Fe(CN)6]4− and KI, DNA competitive binding with ethidium bromide, viscosity measurements, and DNA melting experiments. The results suggest that 1, 2, and 3 complexes bind to CT-DNA through intercalation and follows the order 1 > 2 > 3. Under irradiation at 365 nm, the three complexes have also been found to promote the photocleavage of plasmid pBR322 DNA.  相似文献   

14.
The reductive and the oxidative electron-transfer photochemical reaction system of light-irradiated the mix solutions of Ru(bpy)3 2+ with [Fe(CN)6]4–, [Fe(CN)6]3–, [Fe(CN)5NO]2– and PB (Prussian Blue) have been studied. The double complexes which isolated from the precipitates of the photochemical reaction have been identified by means of Mössbauer spectroscopy. In order to clarify the chemical states of these isolated double complexes, we have (prepared and) studied Mössbauer spectra of the double complexes such as [Ru(bpy)3]3[Fe(CN)6]2.14H2O, [Ru(bpy)3]2[Fe(CN)6].10H2O, [Ru(bpy)3][Fe(CN)5NO].4H2O, and [Ru(bpy)3][PB]2.xH2O.  相似文献   

15.
Theoretical design on a new molecular switch and fluorescent chemosensor double functional device of aza‐crown ether (2,2′‐dipyridine‐embedded N‐(9‐anthraceneyl(pyrenyl)methyl)aza‐15‐crown‐5) was explored. The interactions between ligands and a series of alkaline earth metal cations (Mg2+, Ca2+, Sr2+, and Ba2+) were investigated. The fully optimized geometry structures of the free ligands ( L 1, L 2) and their metal cation complexes ( L 1/M2+, L 2/M2+) were calculated with the B3LYP/6‐31G(d) method. The natural bond orbital analysis, which is based on optimized geometric structures, was used to explore the interaction of L 1/M2+, L 2/M2+ molecules. The absorption spectra of L 1, L 2, L 1/M2+, and L 2/M2+, and their excited states were studied by time‐dependent density functional theory. A new type molecular device L 2(2,2′‐dipyridine‐embedded N‐(9‐pyrenyl methyl)aza‐15‐crown‐5) is designed, which not only has the selectivity for Sr2+, and construct allosteric switch, but also has fluorescent sensor performance.  相似文献   

16.
The Raman spectra of (1 − x)(BMITFSI), xLiTFSI ionic liquids, where 1‐butyl‐3‐methylimidazolium cation (BMI+) and bis(trifluoromethane‐sulfonyl)imide anion (TFSI) are analyzed for LiTFSI mole fractions x < 0.4. As expected from previous studies on similar TFSI‐based systems, most lithium ions are shown to be coordinated within [Li(TFSI)2] anionic clusters. The variation of the self‐diffusion coefficients of the 1H, 19F, and 7Li nuclei, measured by pulsed‐gradient spin‐echo NMR (PGSE‐NMR) as a function of x, can be rationalized in terms of the weighted contribution of BMI+ cations, TFSI ‘free’ anions, and [Li(TFSI)2] anionic clusters. This implies a negative transference number for lithium. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

17.
18.
The nature of the conformational transition of the polymers with Ru (II) polypyridyl complexes covalently attached to poly(acrylic acid) (PAA) and poly(metacrylic acid) (PMAA) has been in studied in aqueous solutions at different pH values. The [PAA-Ru4]8+ and [PMAA-Ru4]8+ polymers has been investigated by means of the luminescence properties of the Ru(bpy)32+ moiety by steady-state and time-resolved luminescence spectroscopy. The pH markedly affects the luminescence spectra and quantum yields of both ruthenium-polyacid complexes in aqueous solution. Another feature investigated in this work was a comparative study of their luminescence quenching by acridinic dyes in solution. The analysis of the kq values obtained indicates that the bimolecular quenching by acridinium and 9-aminoacridinium is more effective in the [PAA-Ru4]8+ complex (6.4×109 and 1.4×109 M−1 s−1, respectively) compared to the [PMAA-Ru4]8+ (2.6×109 and 1.0×109 M−1 s−1). Also, a similar behavior was evidenced for the Ru solely adsorbed onto pure PAA (9.0×109 and 3.4×109 M−1 s−1) and PMAA (1.8×109 and 1.7×109 M−1 s−1) in aqueous solution. The effect of enhancement of quenching rate constant in [PAA-Ru4]8+ system could be ascribed to the higher density of Ru per polymer chain. The average number per chain is similar in both systems, but the molecular weight is lower for [PAA-Ru4]8+. Furthermore, the larger hydrophilic environment provided by the PAA exposes the Ru probe to the outer surface of the polymer in solution.  相似文献   

19.
The gas‐phase elimination kinetics of tetrahydropyranyl phenoxy ethers: 2‐phenoxytetrahydro‐2H‐pyran, 2‐(4‐methoxyphenoxy)tetrahydro‐2H‐pyran, and 2‐(4‐tert‐butylphenoxy)tetrahydro‐2H‐pyran were determined in a static system, with the vessels deactivated with allyl bromide, and in the presence of the free radical inhibitor toluene. The working temperature and pressure were 330 to 390°C and 25 to 89 Torr, respectively. The reactions yielded DHP and the corresponding 4‐substituted phenol. The eliminations are homogeneous, unimolecular, and satisfy a first‐order rate law. The Arrhenius equations for decompositions were found as follows:
  • 2‐phenoxytetrahydro‐2H‐pyran
  • log k1 (s?1) = (14.18 ± 0.21) ? (211.6 ± 0.4) kJ mol?1 (2.303 RT)?1
  • 2‐(4‐methoxyphenoxy)tetrahydro‐2H‐pyran
  • log k1 (s?1) = (14.11 ± 0.18) ? (203.6 ± 0.3) kJ mol?1 (2.303 RT)?1
  • 2‐(4‐tert‐butylphenoxy)tetrahydro‐2H‐pyran
  • log k1 (s?1) = (14.08 ± 0.08) ? (205.9 ± 1.0) kJ mol?1 (2.303 RT)?1
The analysis of kinetic and thermodynamic parameters for thermal elimination of 2‐(4‐substituted‐phenoxy)tetrahydro‐2H‐pyranes suggests that the reaction proceeds via 4‐member cyclic transition state. The results obtained confirm a slight increase of rate constant with increasing electron donating ability groups in the phenoxy ring. The pyran hydrogen abstraction by the oxygen of the phenoxy group appears to be the determinant factor in the reaction rate.  相似文献   

20.
Cu(I) complexes of the type [Cu(L)(PPh3)2]+, where L is the bidentate ligand 4,7‐diphenyl‐1, 10‐phenanthroline (dip) and 3,4,7,8‐tetramethyl‐1,10‐phenanthroline (tem) and their perdeuterated analogues, have been synthesised and the transient resonance Raman spectra of these complexes have been measured. The spectra show two sets of bands, one due to the PPh3 ligands and the other due to L.− created through the metal‐to‐ligand charge transfer transition. Density functional theory calculations have been used to model ligands and complexes in the ground state and good agreement has been found between calculated and measured bands with a mean absolute deviation of 8–10 cm−1 for the ligands and 5 cm−1 for the complexes. Shifts in the bands due to deuteration have also been well predicted, with the shifts for most modes predicted to within 10 cm−1. The structure and spectra of the excited states have been modelled using two approaches. The reduced state [Cu(L.−)(PH3)2] was used for both complexes to predict the changes in the structure of the polypyridyl ligand and for [Cu(dip)(PPh3)2]+ the triplet state was also optimised. Both approaches show that similar structural changes in the ligand are predicted. In the case of [Cu(dip)(PPh3)2]+* and [Cu(dip.−)(PPh3)2], the calculated states are 3A2 and 2A2, respectively, consistent with experiment. Calculations on [Cu(tem)(PPh3)2]+* give a 3B1 state. This is not consistent with experimental results. For [Cu(tem.−)(PPh3)2] both the 2B1 and 2A2 states may be calculated and the experimental spectrum of [Cu(tem)(PPh3)2]+* is closer to that of the 2A2 [Cu(tem.−)(PPh3)2] species. Calculated wavenumbers are compared to measured transient resonance Raman L.− bands and found to have a mean absolute deviation of 8 cm−1 for the triplet state of [Cu(dip)(PPh3)2]+ and 16 cm−1 for the reduced state of [Cu(tem)(PPh3)2]+. Copyright © 2008 John Wiley & Sons, Ltd.  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