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 62 毫秒
1.
The decaaqua-di-rhodium(II) cation has been found to be an interesting starting material in the preparation of dioxygen complexes with different N-donor ligands. Treatment of aqueous HClO4 solution of [Rh2(H2O)10]4+ with NH4OH/NH3, py and/or en results in water exchange and the formation of corresponding [Rh2II(H2O)10−m(base)n(OH)m](4−m)+ derivatives. Reaction of the latter with dioxygen afforded superoxo and/or peroxo complexes, depending on reaction conditions: [Rh2III(O2 −)(NH3)8(OH)2](ClO4)3 (1), [Rh2III(O2 −)(NH3)8(OH)(H2O)](ClO4)4 (2), [Rh2III(O2 2−)(NH3)10](ClO4)4 · 6H2O (3), [Rh2III(O2 −)(py)8(H2O)2](ClO4)5 (4), [Rh2III(O2 2−)(en)4(H2O)2](ClO4)4 (5) and [Rh2III(O2 −)(en)4(H2O)2](ClO4)5 (6). All the obtained complexes were characterized by elemental analysis, mass spectrometry, UV-Vis, IR and ESR spectroscopies and magnetic measurements.  相似文献   

2.
[Rh2Cl2(CO)4] reacts with the ligands L (2-pyridone, 2-thiopyridone, and the isomers 6-methyl-2-thiopyridone, 2-methylmercaptopyridine, and N-methylthiopyridone) to give initially, when L/Rh = 1, the bridged-cleaved compounds cis- [RhCl(CO)2L]. Further additions of 2-methyl- mercaptopyridine, N-methylthiopyridone, or 2-pyridone caused no further change, but 2- thiopyridone and 6-methyl-2-thiopyridone gave new cis-dicarbonyl species (L/Rh = 2) and eventually monocarbonyl species (L/Rh > 3). All these solutions are air-sensitive and air oxidation of a solution of [Rh2Cl2(CO)4] with an excess of 6-methyl-2- thiopyridone gave fac-[Rh(MeC5H3NS)3] the X-ray structure of which shows three equivalent chelating 6-methyl-2-thiopyridonato ligands.  相似文献   

3.
The Schiff base 2,2-bis((4S)-4-benzyl-2-oxazoline) (I) and its coordination complexes with rhodium(I) and palladium(II) (and with 1,5-cyclo-octadiene and allyl ligands) have been characterised by single-crystal X-ray diffraction, mass spectrometry, 13C and 1H NMR spectroscopy: [Rh(C20H20N2O2)(C8H12)][Rh2(C20H20N2O2)2](CF3SO3)3 · (CH3CH2O) (II) and [Pd(C20H20N2O2)(C3H5)]CF3SO3 (III). We discuss the reasons for the formation of two complex cations for Rh(I), [Rh(C20H20N2O2)(C8H12)]+ (IIa) and [Rh2(C20H20N2O2)2]2+ (IIb), and only one for Pd(II).  相似文献   

4.
Complex formation properties of a novel water soluble thiazolyloxime 2-(4-methylthiazol-2-yl)-2-(hydroxyimino)acetic acid (H3L1) with Cu2+ and Ni2+ were investigated in solution by potentiometrical and spectral (UV-Vis, EPR, NMR) methods. All Cu2+ and most of Ni2+ complex species detected in solution were found to have square-planar MN4 core with oxime and heterocyclic nitrogen atoms which was rationalized in terms of destabilizing effect of repulsive interaction between oxygen atom of carboxylic group and nitrogen atom of thiazole ring in N,O-coordinated ligand conformation. It has been found that stability of metal complexes in a series of oxime ligands is dependent upon basicity of nitrogen atom of oxime group. The thiazolyloxime forms less stable complexes with Cu2+ but stronger ones with Ni2+ ions when compared to parent 2-(hydroxyimino)propanoic acid. The lower stability obtained for Cu2+ complexes was elucidated in terms of negative inductive effect of the thiazole and nitrile substituents as well as an effect of intramolecular attractive interaction between thiazolyl sulfur and oxime oxygen atoms in thiazolyloxime. In the case of Ni2+ the complexes formed are square-planar and it is why thiazolyl ligand is more effective in metal ion binding than simple 2-(hydroxyimino)propanoic acid forming only octahedral species. The solid state structure of the Co3+ complex K3[Co(HL1)3]·5.5H2O (1) was studied by X-ray analysis. The thiazolyloxime ligand is coordinated to Co3+ via oxime nitrogen and carboxylate oxygen atoms forming five-membered chelate rings.  相似文献   

5.
Pentamethylcyclopentadienyl rhodium bipyridine ([Cp*Rh(bpy)(H2O)]2+) is a versatile catalyst to promote biocatalytic redox reactions. However, its major drawback lies in the mutual inactivation of [Cp*Rh(bpy)(H2O)]2+ and the biocatalyst. This interaction was investigated using the alcohol dehydrogenase from Thermus sp. ATN1 (TADH) as model enzyme. TADH binds 4 equiv. of [Cp*Rh(bpy)(H2O)]2+ without detectable decrease in catalytic activity and stability. Higher molar ratios lead to time-, temperature-, and concentration-dependent inactivation of the enzyme suggesting [Cp*Rh(bpy)(H2O)]2+ to function as an ‘unfolding catalyst’. This detrimental activity can be circumvented using strongly coordinating buffers (e.g. (NH4)2SO4) while preserving its activity as NAD(P)H regeneration catalyst under electrochemical reaction conditions.  相似文献   

6.
The kinetic inertness of Rh(III) has facilitated investigations of the early stages of hydrolytic polymerization of [Rh(OH2)6]3+. These studies have led to the synthesis, solution characterization and determination of the X-ray crystal structure of a series of polynuclear aqua ions, which range in nuclearity from binuclear to tetranuclear. The inertness of Rh(III) has enabled the identification of three structural forms of the trinuclear aqua ion which vary in their degree of condensation as well as detailed kinetic and mechanistic studies of water exchange on the binuclear aqua ion, [Rh2(μ-OH)2(OH2)8]4+, applying both 18O isotopic labeling and 17O NMR spectroscopy. Strategies have been developed which led to the quite efficient synthesis of heteropolynuclear Rh(III)-Cr(III) aqua ions and a heterobinuclear Rh(III)-Ir(III) aqua ion. Elucidation of the composition, properties and structure of heteropolynuclear aqua ions has been possible by applying techniques that had been instrumental in the development of the chemistry of homometallic aqua ions. Kinetic and thermodynamic studies of the Rh(III)-Cr(III) aqua ions generally revealed thermodynamic and activation parameters that corresponded to those for similar processes occurring at Cr(III), as may have been expected given the greater lability of Cr(III) relative to Rh(III). Inductive effects caused by the Rh(III) centers do, however, influence reactivity as exemplified by the slower than expected rate of cleavage of [CrRh2(μ-OH)4(OH2)10]5+.  相似文献   

7.
《Luminescence》2003,18(1):49-57
The chemiluminescence reaction of lucigenin (Luc2+?2NO3?, N,N′‐dimethyl‐9,9′‐biacridinium dinitrate) at gold electrodes in dioxygen‐saturated alkaline aqueous solutions (pH 10) was investigated in detail by the use of electrochemical emission spectroscopy. We noted that both O2 and Luc2+ are reduced on a gold electrode in aqueous solution of pH 10 in almost the same potential region. From this fact, we expected chemiluminescence based on a radical–radical coupling reaction of superoxide ion (O2·?) and one‐electron reduced form of Luc2+ (Luc·+, a radical cation). Chemiluminescence was actually observed in the potential range where O2 and Luc2+ were simultaneously reduced at the electrodes. The effects were examined upon addition of enzymes, i.e. superoxide dismutase (SOD) and catalase, into the solution and the substitution of heavy water (D2O) for light water (H2O) as a solvent on the chemiluminescence. In the presence of native and active SOD, chemiluminescence was completely absent. On the other hand, chemiluminescence was observed, unchanged in the presence of either denatured and inert SOD or catalase. In addition, the amount of chemiluminescence in D2O solution was about three times greater than that in H2O solution. These results, together with cyclic voltammetric results, suggest that O2·? participates directly in the chemiluminescence but H2O2 does not, and the chemiluminescence results from the coupling reaction between O2·? and Luc·+ under the present experimental conditions. These chemically unstable species, O2·? and Luc·+, are produced during the simultaneous electroreduction of O2 and Luc2+. The coupling reaction between those radical species would lead to the formation of a dioxetane‐type intermediate and, finally, to chemiluminescence. The chemiluminescence reaction mechanism is discussed. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

8.
A multiphase model of metal ion speciation in human interstitial fluid was constructed and the effect of Pr(III) on Ca(II) speciation was studied. Results show that free Ca2+, [Ca(HCO3)], and [Ca(Lac)] are the main species of Ca(II). Because of the competition of Pr(III) for ligands with Ca(II), the percentages of free Ca2+, [Ca(Lac)], and [Ca(His)(Thr)H3] increase gradually and the percentages of CaHPO4(aq) and [Ca(Cit)(His)H2] decrease gradually with the increase in the total concentration of Pr(III). However, the percentages of [Ca(HCO3)] and CaCO3(aq) first increase and then begin to decrease when the total concentration of Pr(III) exceeds 6.070×10−4 M.  相似文献   

9.
The reaction of [Rh2(acam)4(H2O)2]ClO4 (1) (Hacam = acetamide) with K2PtCl4 in aqueous solution gave crystals of [Rh2(acam)4(H2O)2][Rh2(acam)4{(μ-Cl)2PtCl2}] · 2H2O (2). The reaction of 1 with K2PdCl4 produced the palladium analog [Rh2(acam)4(H2O)2][Rh2(acam)4{(μ-Cl)2PdCl2}] · 2H2O (3) and a small amount of an aquated palladium complex [Rh2(acam)4{(μ-Cl)2PdCl(H2O)}] · H2O (4). Complexes 2 and 3 have anionic chains of [Rh2(acam)4{(μ-Cl)2MCl2}] (M = Pt, Pd), while 4 includes neutral chains of [Rh2(acam)4{(μ-Cl)2PdCl(H2O)}]. Although all of the structures include infinite chains of (-Rh-Rh-Cl-M-Cl-)n (M = Pt, Pd), the chain structures are different; zigzag for 2 and 3 and helical for 4. In the structures of 2 and 3, the counter cation [Rh2(acam)4(H2O)2]+ made a hydrogen-bonded chain with the crystallization water molecules. The cationic chains and the anionic chains are connected with hydrogen bonds. In the structure of 4, the chains are also linked together by direct hydrogen bonds between the chains and those with the crystallization water molecules. ESR spectra of the powdered samples of 2 and 3 at 77 K were consistent with a rhombic structure: for 2, g1 = 2.111, g2 = 2.054, g3 = 2.004; for 3, g1 = 2.115, g2 = 2.057, g3 = 2.007. These results indicate that there is a spin flip-flop exchange between the cations, [Rh2(acam)4(H2O)2]+, and the units in the anionic chains. The electrical conductivities of 2 and 3 were in the order of 10−7 S cm−1 at room temperature.  相似文献   

10.
Identity of the rhizotoxic aluminium species   总被引:11,自引:3,他引:8  
The aluminium (III) released from soil minerals to the soil solution under acid conditions may appear as hexaaquaaluminium (Al(H2O)6 3+, or Al3+ for convenience) or may react with available ligands to form additional chemical species. That one or more of these species is rhizotoxic (inhibitory to root elongation) has been known for many decades, but the identity of the toxic species remains problematical for the following reasons. 1. Several Al species coexist in solution so individual species cannot be investigated in isolation, even in artificial culture media. 2. The activities of individual species must be calculated from equilibrium data that may be uncertain. 3. The unexpected or undetected appearance of the extremely toxic triskaidekaaluminium (AlO4Al12(OH)24(H2O)12 7+ or Al13) may cause misatribution of toxicity to other species, especially to mononuclear hydroxy-Al. 4. If H+ ameliorates Al3+ toxicity, or vice versa, then mononuclear hydroxy-Al may appear to be toxic when it is not. 5. The identity and activities of the Al species contacting the cell surfaces are uncertain because of the H+ currents through the root surface and because of surface charges. This article considers the implications of these problems for good experimental designs and critically evaluates current information regarding the relative toxicities of selected Al species. It is concluded that polycationic Al (charge >2) is rhizotoxic as are other polyvalent cations.  相似文献   

11.
《Inorganica chimica acta》1986,119(2):227-232
Interaction between D-glucuronic acid and hydrated uranyl salts has been studied in aqueous solution and solid complexes of the type UO2(D- glucuronate)X·2H20 and UO2(D-glucuronate)2·2H2O, where X = CI, Br or NO3, are isolated and characterized by means of FT-IR and proton-NMR spectroscopy.On comparison with the structurally identified Ca(D-glucuronate)Br·3H2O compound, it is concluded that the UO22+ cation binds to two D- glucuronate moieties in uranylsugar complexes via O6, O5 oxygen atoms (ionized carboxyl group) of the first and O6′, 04 (non-ionized carboxyl group) of the second sugar moiety, whereas in the UO2(D- glucuronate)2·2H2O salt the uranyl ion is bonded to two sugar anions through O6, O6′ oxygen atoms of the ionized carboxyl group, resulting in a six- coordination geometry around the uranium ion. The strong intermolecular hydrogen bonding network of the free acid is rearranged upon sugar metalation and the sugar moiety showed β-anomer conformation both in the free acid and in these uranylsugar complexes.  相似文献   

12.
The reaction between [Rh(H2O)6](ClO4)3 and the monoanion Hdopn (H2dopn=bis(diacetylmonoxime-imino)propane 1,3=3,9-dimethyl-4,8-diazaundeca-3,8-diene-2,10-dione dioxime) afforded a new dimeric rhodium(II) compound of formula [Rh(Hdopn)(H2O)]2(ClO4)2 · H2O (1). Treatment of methanolic solution of 1 with NaX (X=Cl, Br, I) results in the replacement of water with halides in 1, leading to the formation of [Rh(Hdopn)X]2 rhodium(II) dimers. The X-ray crystal structure of [Rh(Hdopn)Cl]2 · 0.5H2O (2) was determined showing a [Rh(II)-Rh(II)] core. Upon the reaction of 1 with NaI carried out in air, [Rh(Hdopn)(I)2] (3) was isolated and characterized by a single-crystal X-ray diffraction analysis.  相似文献   

13.
[Rh2(μ-Cl)2(cod)2] reacts with Ph2PCH2CH2OMe (PC2O), Ph2P(CH2)3NMe2 (PC3N), PBunPh2 or PPh3 to give [Rh(cod)L2]Cl (L = PC2O, PC3N, PBunPh2, PPh3). In the presence of hydrogen, [Rh(cod)L2]Cl is converted to [RhClH2L3]. In contrast, [Rh(cod)(PC2O)2]BPh4 reacts with H2 to give [RhH2(PC2O)2S2]BPh4 (S = solvent). With Ph2PCH2CH2NMe2 (PC2N) or Ph2PCH2CH2SMe (PC2S), [Rh2(μ-Cl)2(cod)2] reacts to form the chelate complexes cis- [Rh(PC2N)2]+ or cis-[Rh(PC2S)2]+, neither of which reacts with hydrogen under ambient conditions. The products of the reactions are characterized in situ by 31P1H NMR spectroscopy.  相似文献   

14.
Structures of rhodium(II) binuclear complexes [Rh2(OOCCH3)2(bpy)2(H2O){(CH3)2CHOH}][B(C6H5)4]2 · H2O (1), [Rh2Cl2(OOCCH3)2(bpy)2] · 2H2O (2), [Rh2Br2(OOCCH3)2(bpy)2] · 3H2O (3), and [Rh2I2(OOCCH3)2(bpy)2] (4), as well as an unprecedented wire with infinite Rh-Rh chain, {[Rh4(μ-OOCH)4(bpy)4](BF4)}n · 0.5nC4H8O2 (5), have been determined and discussed. Mass spectra of complexes [Rh2(OOCMe)2(bpy)2(H2O)2](MeCOO)2 and [Rh2(OOCMe)2(phen)2(H2O)2](MeCOO)2 have showed stability of polynuclear cations with rhodium in oxidation states in the range +1.25 to +1.75.  相似文献   

15.
In this work, porous monolayer nickel‐iron layered double hydroxide (PM‐LDH) nanosheets with a lateral size of ≈30 nm and a thickness of ≈0.8 nm are successfully synthesized by a facile one‐step strategy. Briefly, an aqueous solution containing Ni2+ and Fe3+ is added dropwise to an aqueous formamide solution at 80 °C and pH 10, with the PM‐LDH product formed within only 10 min. This fast synthetic strategy introduces an abundance of pores in the monolayer NiFe‐LDH nanosheets, resulting in PM‐LDH containing high concentration of oxygen and cation vacancies, as is confirmed by extended X‐ray absorption fine structure and electron spin resonance measurements. The oxygen and cation vacancies in PM‐LDH act synergistically to increase the electropositivity of the LDH nanosheets, while also enhancing H2O adsorption and bonding strength of the OH* intermediate formed during water electrooxidation, endowing PM‐LDH with outstanding performance for the oxygen evolution reaction (OER). PM‐LDH offers a very low overpotential (230 mV) for OER at a current density of 10 mA cm?2, with a Tafel slope of only 47 mV dec?1, representing one of the best OER performance yet reported for a NiFe‐LDH system. The results encourage the wider utilization of porous monolayer LDH nanosheets in electrocatalysis, catalysis, and solar cells.  相似文献   

16.
Abstract

The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5′-mono- and -triphosphates in D2O solution at Ph′=3 were measured. The paramagnetic probes were [Cr(III)(H2O) 6]3+, [Cr(III)(H2O)3 (HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III) (H2O)3(UTP)?, while the matrix nucleotides (0.1 M) were H2AMP, HIMP?, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]?/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(α,β)], [Cr(III)(NH3)4(HPP)(α,β)], [Co(III)(NH3)3(H2PPP)(α,βγ)] and [Co(III)(NH3)4(HPP)(α,β)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(α,βγ)] complexes as well as the structures of the outer sphere [Cr(III) (H2O)6]3+-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)? species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(α,βγ)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7) (O…N = 2.82 Å). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2′)H group and the adenine N(3) (O…N = 2.80 Å) as well as phosphate oxygen atoms and a water molecule (O…O = 2.47 Å). The metal center has an almost regular octahedral coordination geometry.

The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar “anti” conformation around the N(9)-C(l′) glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)? complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O…O = 2.63 Å O…N = 2.72, 2.70 Å). For the H2AMP complex, the [Cr(III) (H2O)c]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s?1 (H(2)) for HIMP?. These results suggest that the dynamic relaxation effects can be only partially understood with molecular mechanics computations, although the success of the geometry calculations suggests that future efforts in the development of computational methods are justified.  相似文献   

17.
《Inorganica chimica acta》1988,153(3):155-159
The interaction of D-glucose with hydrated uranyl salts has been investigated in solution and solid adducts of the type UO2(D-glucose)X2·2H2O, where X = Cl, Br, NO3 and 0.5 SO42− have been isolated. These adducts are characterized by means of FT-IR, 1H NMR and molar conductivity measurements.Spectroscopic evidence suggested that UO22+ cation could be bonded to one D-glucose molecule (possibly through O(1)H and O(2)H hydroxyl groups) and to two H2O, resulting in six-coordination around the uranium ion.The strong sugar H-bonding network is perturbed, on metal ion interaction and the D-glucose α-anomeric structure is favoured, upon uranyl cation coordination.  相似文献   

18.
The S-bridged trinuclear complexes composed of heavy d6 metal ions, [RhIII{M(aet)3}2]3+ (M=IrIII(1), RhIII(2); aet = 2-aminoethanethiolate), have been prepared by the reactions of fac(S)-[M(aet)3] with RhCl3 · 3H2O. The complexes were separated into meso (1a, 2a) and rac (1b, 2b) isomers by SP-Sephadex C-25 column chromatography. 1b and 2b were optically resolved by the column chromatographic method and characterized by CD spectroscopy. Crystal structures of 1a, 1b and 2a were determined by X-ray diffraction, and it was found that they consist of linear-type trinuclear structures. The central Rh(III) ion in the present complexes has d6 electronic configuration with the non-degenerated A-type cubic field term, and showed long Rh?M distances, acute S-M-S angles and obtuse Rh-S-M angles. These are in contrast with the complexes having the degenerated T-type cubic field term such as [M{M(aet)3}2]n+ (M=VIII, MoIV and ReIII, M=IrIII, RhIII, n=3 or 4). All the isomers have been comparatively characterized and discussed in solid state and the solution for spectrochemical and electrochemical properties.  相似文献   

19.
From solutions of seleno bridged triangular cluster Mo3Se4(aq)4+ in HCl, crystalline adducts with cucurbituril (Cuc, C36H36N24O12) of different composition, depending on HCl concentration, were isolated. From 2 M HCl, a monosubstituted cationic cluster crystallizes as {[Mo3Se4Cl(H2O)8]2(C36H36N24O12)}Cl6·16H2O (1). Increase in HCl concentration to 6 M gives a pentasubstituted anionic species, (H3O)2[Mo3Se4Cl5(H2O)4]2(C36H36N24O12)·15H2O (2). The crystal structures of 1 and 2 were determined by X-ray crystallography. Each portal of Cuc in 1 is covered with cluster cations [Mo3Se4Cl(H2O)8]3+ like a ‘lid’ on a ‘barrel’. Six water molecules in the trans position to the core μ2-Se form complementary hydrogen bonds with oxygen atoms of Cuc (O?O, 2.713-3.067 Å). In 2 the complementarity is lost and the main structure building factor is short Se?Se interactions (Se?Se, 2.96-3.43 Å) between two adjacent anionic clusters. Stereochemistry of halide substitution in the triangular clusters M3Q4 is analyzed.  相似文献   

20.
A mutiphase model of metal ion speciation in human interstitial fluid was constructed and the effect of Pr(III) on Ca(II) speciation was studied. Results show that Ca(II) mainly distributes in free Ca2+, [Ca(HCO3)], and [Ca(Lac)]. Because of the competition of Pr(III) for ligands with Ca(II), with the total concentration of Pr(III) rising, the percentages of free Ca2+, [Ca(Lac)] and [Ca(His)(Thr)H3], gradually increase and the percentages of CaHPO4(aq) and [Ca(Cit)(His)H2] gradually decrease. However, the percentages of [Ca(HCO3)] and CaCO3(aq) first increase, and then begin to decrease when the total concentration of Pr(III) exceeds 6.070×10−4 M.  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