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.
C2‐C70(CF3)8 was found to be a very promising substrate in the Bingel and the Bingel–Hirsch reactions combining perfect regioselectivity with much higher reactivity compared to its analogs. The reactions with diethyl malonate yield a single isomer of the monoadduct C70(CF3)8[C(CO2Et)2] and a single C2‐symmetrical bisadduct C70(CF3)8[C(CO2Et)2]2. The Bingel–Hirsch variation is particularly interesting in that it additionally affords, in a similar regioselective manner, the unexpected alkylated derivatives C70(CF3)8[CH(CO2Et)2]H and C70(CF3)8[C(CO2Et)2][CH(CO2Et)2]H. The novel compounds have been isolated and structurally characterized by means of 1H and 19F NMR spectroscopy as well as single‐crystal X‐ray diffraction. The mechanistic and regiochemical aspects of the reaction are explained with the aid of DFT calculations.  相似文献   

2.
4‐Fluorinated levoglucosans were synthesised to test if OH???F H‐bonds are feasible even when the O???F distance is increased. The fluorinated 1,6‐anhydro‐β‐D ‐glucopyranoses were synthesised from 1,6 : 3,4‐dianhydro‐β‐D ‐galactopyranose ( 8 ). Treatment of 8 with KHF2 and KF gave 43% of 4‐deoxy‐4‐fluorolevoglucosan ( 9 ), which was transformed into the 3‐O‐protected derivatives 13 by silylation and 15 by silylation, acetylation, and desilylation. 4‐Deoxy‐4‐methyllevoglucosan ( 19 ) and 4‐deoxylevoglucosan ( 21 ) were prepared as reference compounds that can only form a bivalent H‐bond from HO? C(2) to O? C(5). They were synthesised from the iPr3Si‐protected derivative of 8 . Intramolecular bifurcated H‐bonds from HO? C(2) to F? C(4) and O? C(5) of the 4‐fluorinated levoglucosans in CDCl3 solution are evidenced by the 1H‐NMR scalar couplings h1J(F,OH) and 3J(H,OH). The OH???F H‐bond over an O???F distance of ca. 3.0 Å is thus formed in apolar solvents, at least when favoured by the simultaneous formation of an OH???O H‐bond.  相似文献   

3.
Two novel intrinsically microporous copolyimides synthesized by condensation reaction of 4,4′‐(hexafluoroisopropylidene) diphthalic anhydride (6FDA), 3,3,3′,3′‐tetramethyl‐1,1′‐spirobisindane‐5,5′‐diamino‐6,6′‐diol, and 3,5‐diaminobenzoic acid with diamine ratios of 80/20 (Co‐80/20) and 50/50 (Co‐50/50) are reported. Unexpectedly, the Co‐80/20 not only demonstrates higher microporosity (300 m2 g−1) than the PIM‐6FDA‐OH homopolymer (190 m2 g−1) but also exhibits simultaneously enhanced CO2 permeability (from 119 to 171 Barrer) and CO2/CH4 selectivity (from 35 to 41) after thermal annealing at 250 °C. This higher permeability originates from enhanced diffusivity (D CO2) and the higher selectivity results from its increased diffusion selectivity (D CO2/D CH4). After crosslinking at 300 °C, the Co‐80/20 exhibits an even higher CO2 permeability (261 Barrer) and almost unchanged CO2/CH4 selectivity.

  相似文献   


4.
Nucleophilic incorporation of [18F]F? under aqueous conditions holds several advantages in radiopharmaceutical development, especially with the advent of complex biological pharmacophores. Sulfonyl fluorides can be prepared in water at room temperature, yet they have not been assayed as a potential means to 18F‐labelled biomarkers for PET chemistry. We developed a general route to prepare bifunctional 4‐formyl‐, 3‐formyl‐, 4‐maleimido‐ and 4‐oxylalkynl‐arylsulfonyl [18F]fluorides from their sulfonyl chloride analogues in 1:1 mixtures of acetonitrile, THF, or tBuOH and Cs[18F]F/Cs2CO3(aq.) in a reaction time of 15 min at room temperature. With the exception of 4‐N‐maleimide‐benzenesulfonyl fluoride ( 3 ), pyridine could be used to simplify radiotracer purification by selectively degrading the precursor without significantly affecting observed yields. The addition of pyridine at the start of [18F]fluorination (1:1:0.8 tBuOH/Cs2CO3(aq.)/pyridine) did not negatively affect yields of 3‐formyl‐2,4,6‐trimethylbenzenesulfonyl [18F]fluoride ( 2 ) and dramatically improved the yields of 4‐(prop‐2‐ynyloxy)benzenesulfonyl [18F]fluoride ( 4 ). The N‐arylsulfonyl‐4‐dimethylaminopyridinium derivative of 4 ( 14 ) can be prepared and incorporates 18F efficiently in solutions of 100 % aqueous Cs2CO3 (10 mg mL?1). As proof‐of‐principle, [18F] 2 was synthesised in a preparative fashion [88(±8) % decay corrected (n=6) from start‐of‐synthesis] and used to radioactively label an oxyamino‐modified bombesin(6–14) analogue [35(±6) % decay corrected (n=4) from start‐of‐synthesis]. Total preparation time was 105–109 min from start‐of‐synthesis. Although the 18F‐peptide exhibited evidence of proteolytic defluorination and modification, our study is the first step in developing an aqueous, room temperature 18F labelling strategy.  相似文献   

5.
[(ArPMI)Mo(CO)4] complexes (PMI=pyridine monoimine; Ar=Ph, 2,6‐di‐iso‐propylphenyl) were synthesized and their electrochemical properties were probed with cyclic voltammetry and infrared spectroelectrochemistry (IR‐SEC). The complexes undergo a reduction at more positive potentials than the related [(bipyridine)Mo(CO)4] complex, which is ligand based according to IR‐SEC and DFT data. To probe the reaction product in more detail, stoichiometric chemical reduction and subsequent treatment with CO2 resulted in the formation of a new product that is assigned as a ligand‐bound carboxylate, [(PMI)Mo(CO)3(CO2)]2?, by NMR spectroscopic methods. The CO2 adduct [(PMI)Mo(CO)3(CO2)]2? could not be isolated and fully characterized. However, the C?C coupling between the CO2 molecule and the PDI ligand was confirmed by X‐ray crystallographic characterization of one of the decomposition products of [(PMI)Mo(CO)3(CO2)]2?.  相似文献   

6.
The reaction of propargylic amines and CO2 can provide high‐value‐added chemical products. However, most of catalysts in such reactions employ noble metals to obtain high yield, and it is important to seek eco‐friendly noble‐metal‐free MOFs catalysts. Here, a giant and lantern‐like [Zn116] nanocage in zinc‐tetrazole 3D framework [Zn22(Trz)8(OH)12(H2O)9?8 H2O]n Trz=(C4N12O)4? ( 1 ) was obtained and structurally characterized. It consists of six [Zn14O21] clusters and eight [Zn4O4] clusters. To our knowledge, this is the highest‐nuclearity nanocages constructed by Zn‐clusters as building blocks to date. Importantly, catalytic investigations reveal that 1 can efficiently catalyze the cycloaddition of propargylic amines with CO2, exclusively affording various 2‐oxazolidinones under mild conditions. It is the first eco‐friendly noble‐metal‐free MOFs catalyst for the cyclization of propargylic amines with CO2. DFT calculations uncover that ZnII ions can efficiently activate both C≡C bonds of propargylic amines and CO2 by coordination interaction. NMR and FTIR spectroscopy further prove that Zn‐clusters play an important role in activating C≡C bonds of propargylic amines. Furthermore, the electronic properties of related reactants, intermediates and products can help to understand the basic reaction mechanism and crucial role of catalyst 1 .  相似文献   

7.
To enable a comparison between a C—H…X hydrogen bond and a halogen bond, the structures of two fluorous‐substituted pyridinium iodide salts have been determined. 4‐[(2,2‐Difluoroethoxy)methyl]pyridinium iodide, C8H10F2NO+·I, (1), has a –CH2OCH2CF2H substituent at the para position of the pyridinium ring and 4‐[(3‐chloro‐2,2,3,3‐tetrafluoropropoxy)methyl]pyridinium iodide, C9H9ClF4NO+·I, (2), has a –CH2OCH2CF2CF2Cl substituent at the para position of the pyridinium ring. In salt (1), the iodide anion is involved in one N—H…I and three C—H…I hydrogen bonds, which, together with C—H…F hydrogen bonds, link the cations and anions into a three‐dimensional network. For salt (2), the iodide anion is involved in one N—H…I hydrogen bond, two C—H…I hydrogen bonds and one C—Cl…I halogen bond; additional C—H…F and C—F…F interactions link the cations and anions into a three‐dimensional arrangement.  相似文献   

8.
An innovative technique to obtain high‐surface‐area mesostructured carbon (2545 m2 g?1) with significant microporosity uses Teflon as the silica template removal agent. This method not only shortens synthesis time by combining silica removal and carbonization in a single step, but also assists in ultrafast removal of the template (in 10 min) with complete elimination of toxic HF usage. The obtained carbon material (JNC‐1) displays excellent CO2 capture ability (ca. 26.2 wt % at 0 °C under 0.88 bar CO2 pressure), which is twice that of CMK‐3 obtained by the HF etching method (13.0 wt %). JNC‐1 demonstrated higher H2 adsorption capacity (2.8 wt %) compared to CMK‐3 (1.2 wt %) at ?196 °C under 1.0 bar H2 pressure. The bimodal pore architecture of JNC‐1 led to superior supercapacitor performance, with a specific capacitance of 292 F g?1 and 182 F g?1 at a drain rate of 1 A g?1 and 50 A g?1, respectively, in 1 m H2SO4 compared to CMK‐3 and activated carbon.  相似文献   

9.
Conjugated microporous polymers (CMPs), with the virtue of high porosity and optoelectronic activity, are attracting increasing research interest and have been used in various environmental and energy areas. Efficient synthesis and the exploitation of new functionalities are the research hotspots in the CMPs research area. Covalent triazine frameworks (CTFs) synthesized by CF3SO3H catalyzed trimerization reactions show properties quite alike to CMPs and this method avoids the use of noble metal catalysts. In this study, a series of novel fluorescent covalent triazine‐based frameworks (F‐CTFs) is prepared using different tetra‐cyano compounds as the starting monomers. Both porosity and fluorescence properties of the F‐CTFs can be adjusted by the monomer structure. Gas adsorption measurement reveals that F‐CTF1 with the largest surface area of 896 m2 g−1 shows the highest CO2 uptake of 3.29 mmol g−1 at 273 K and 1.13 bar among the polymers. Taking advantages of their large surface areas and strong fluorescence, these F‐CTFs could be used as efficient chemical sensing agents for various nitroaromatic compounds as well.

  相似文献   


10.
The first examples of magnesium(I) dimers bearing tripodal ligands, [(Mg{κ3N,N′,O‐(ArNCMe)2(OCCPh2)CH})2] [Ar=2,6‐iPr2C6H3 (Dip) 7 , 2,6‐Et2C6H3 (Dep) 8 , or mesityl (Mes) 9 ] have been prepared by post‐synthetic modification of the β‐diketiminato ligands of previously reported magnesium(I) systems, using diphenylketene, O?C?CPh2. In contrast, related reactions between β‐diketiminato magnesium(I) dimers and the isoelectronic ketenimine, MesN?C?CPh2, resulted in reductive insertion of the substrate into the Mg?Mg bond of the magnesium(I) reactant, and formation of [{(Nacnac)Mg}2{μ‐κ2N,C‐(Mes)NCCPh2}] (Nacnac=[(ArNCMe)2CH]?; Ar=Dep 10 or Mes 11 ). Reactions of the four‐coordinate magnesium(I) dimer 8 with excess CO2 are readily controlled, and cleanly give carbonate [(LMg)2(μ‐κ22‐CO3)] 12 (L=[κ3N,N′,O‐(DepNCMe)2(OCCPh2)CH]?; thermodynamic product), or oxalate [(LMg)2(μ‐κ22‐C2O4)] 13 (kinetic product), depending on the reaction temperature. Compound 12 and CO are formed by reductive disproportionation of CO2, whereas 13 results from reductive coupling of two molecules of the gas. Treatment of 8 with an excess of N2O cleanly gives the μ‐oxo complex [(LMg)2(μ‐O)] 14 , which reacts facilely with CO2 to give 12 . This result presents the possibility that 14 is an intermediate in the formation of 12 from the reaction of 8 and CO2. In contrast to its reactions with CO2, 8 reacts with SO2 over a wide temperature range to give only one product; the first example of a magnesium dithionite complex, [(LMg)2(μ‐κ22‐S2O4)] 16 , which is formed by reductive coupling of two molecules of SO2, and is closely related to f‐block metal dithionite complexes derived from similar SO2 reductive coupling processes. On the whole, this study strengthens previously proposed analogies between the reactivities of magnesium(I) systems and low‐valent f‐block metal complexes, especially with respect to small molecule activations.  相似文献   

11.
Guanidines and amidines prove to be highly efficient metal‐free catalysts for the reduction of CO2 to methanol with hydroboranes such as 9‐borabicyclo[3.3.1]nonane (9‐BBN) and catecholborane (catBH). Nitrogen bases, such as 1,5,7‐triazabicyclo[4.4.0]dec‐5‐ene (TBD), 7‐methyl‐1,5,7‐triazabicyclo[4.4.0]dec‐5‐ene (Me‐TBD), and 1,8‐diazabicycloundec‐7‐ene (DBU), are active catalysts for this transformation and Me‐TBD can catalyze the reduction of CO2 to methoxyborane at room temperature with TONs and TOFs of up to 648 and 33 h?1 (25 °C), respectively. Formate HCOOBR2 and acetal H2C(OBR2)2 derivatives have been identified as reaction intermediates in the reduction of CO2 with R2BH, and the first C?H‐bond formation is rate determining. Experimental and computational investigations show that TBD and Me‐TBD follow distinct mechanisms. The N?H bond of TBD is reactive toward dehydrocoupling with 9‐BBN and affords a novel frustrated Lewis pair (FLP) that can activate a CO2 molecule and form the stable adduct 2 , which is the catalytically active species and can facilitate the hydride transfer from the boron to the carbon atoms. In contrast, Me‐TBD promotes the reduction of CO2 through the activation of the hydroborane reagent. Detailed DFT calculations have shown that the computed energy barriers for the two mechanisms are consistent with the experimental findings and account for the reactivity of the different boron reductants.  相似文献   

12.
We report a family of cationic lead halide layered materials, formulated as [Pb2X2]2+[O2C(CH)2CO2] (X=F, Cl, Br), exhibiting pronounced broadband white‐light emission in bulk form. These well‐defined PbX‐based structures achieve an external quantum efficiency as high as 11.8 %, which is comparable to the highest reported value (ca.9 %) for broadband phosphors based on layered organolead halide perovskites. More importantly, our cationic materials are ultrastable lead halide materials, which overcome the air/moisture‐sensitivity problems of lead perovskites. In contrast to the perovskites and other bulk emitters, the white‐light emission intensity of our materials remains undiminished after continuous UV irradiation for 30 days under atmospheric conditions (ca.60 % relative humidity). Our mechanistic studies confirm that the broadband emission is ascribed to short‐range electron‐phonon coupling in the strongly deformable lattice and generated self‐trapped carriers.  相似文献   

13.
Electrochemical CO2 reduction relies on the availability of highly efficient and selective catalysts. Herein, we report a general strategy to boost the activity of metal–organic frameworks (MOFs) towards CO2 reduction via ligand doping. A strong electron‐donating molecule of 1,10‐phenanthroline was doped into Zn‐based MOFs of zeolitic imidazolate framework‐8 (ZIF‐8) as CO2 reduction electrocatalyst. Experimental and theoretical evidences reveal that the electron‐donating nature of phenanthroline enables a charge transfer, which induces adjacent active sites at the sp2 C atoms in the imidazole ligand possessing more electrons, and facilitates the generation of *COOH, hence leading to improved activity and Faradaic efficiency towards CO production.  相似文献   

14.
We report a family of cationic lead halide layered materials, formulated as [Pb2X2]2+[O2C(CH)2CO2] (X=F, Cl, Br), exhibiting pronounced broadband white‐light emission in bulk form. These well‐defined PbX‐based structures achieve an external quantum efficiency as high as 11.8 %, which is comparable to the highest reported value (ca.9 %) for broadband phosphors based on layered organolead halide perovskites. More importantly, our cationic materials are ultrastable lead halide materials, which overcome the air/moisture‐sensitivity problems of lead perovskites. In contrast to the perovskites and other bulk emitters, the white‐light emission intensity of our materials remains undiminished after continuous UV irradiation for 30 days under atmospheric conditions (ca.60 % relative humidity). Our mechanistic studies confirm that the broadband emission is ascribed to short‐range electron‐phonon coupling in the strongly deformable lattice and generated self‐trapped carriers.  相似文献   

15.
A new porous organic polymer, SNU‐C1 , incorporating two different CO2‐attracting groups, namely, carboxy and triazole groups, has been synthesized. By activating SNU‐C1 with two different methods, vacuum drying and supercritical‐CO2 treatment, the guest‐free phases, SNU‐C1‐va and SNU‐C1‐sca , respectively, were obtained. Brunauer–Emmett–Teller (BET) surface areas of SNU‐C1‐va and SNU‐C1‐sca are 595 and 830 m2g?1, respectively, as estimated by the N2‐adsorption isotherms at 77 K. At 298 K and 1 atm, SNU‐C1‐va and SNU‐C1‐sca show high CO2 uptakes, 2.31 mmol g?1 and 3.14 mmol g?1, respectively, the high level being due to the presence of abundant polar groups (carboxy and triazole) exposed on the pore surfaces. Five separation parameters for flue gas and landfill gas in vacuum‐swing adsorption were calculated from single‐component gas‐sorption isotherms by using the ideal adsorbed solution theory (IAST). The data reveal excellent CO2‐separation abilities of SNU‐C1‐va and SNU‐C1‐sca , namely high CO2‐uptake capacity, high selectivity, and high regenerability. The gas‐cycling experiments for the materials and the water‐treated samples, experiments that involved treating the samples with a CO2‐N2 gas mixture (15:85, v/v) followed by a pure N2 purge, further verified the high regenerability and water stability. The results suggest that these materials have great potential applications in CO2 separation.  相似文献   

16.
Reaction of bisalkylidyne cluster compounds [Fe3(CO)93‐CR)2] ( 1a—d ) ( a , R = H; b , R = F; c , R = Cl; d , R = Br) with the phosphaalkyne t‐C4H9‐C≡P ( 2 ) yield a single isomer of the phosphaferrole cluster [Fe3(CO)8][CR‐C(t‐Bu)‐P‐CR] ( 3a—d ). However, the three isomeric compounds [Fe3(CO)8][C(OEt)‐C(t‐Bu)‐P‐C(Me)] ( 5a ), [Fe3(CO)8][C(Me)‐C(t‐Bu)‐P‐C(OEt)] ( 5b ), and [Fe3(CO)8][C(OEt)‐C(Me)‐C(t‐Bu)‐P] ( 5c ) are obtained in the reaction of [Fe3(CO)93‐CMe)(μ3‐C‐OEt)] ( 4 ) with 2 . As the phosphaferroles 3 possess a lone pair of electrons at the phosphorus atom they can act as ligands. [Fe3(CO)8][CF‐C(t‐Bu)‐P‐CF]MLn ( 7a—c ) ( a , MLn = Cr(CO)5; b , MLn = CpMn(CO)2; c , MLn = Cp*Mn(CO)2) were formed from 3b and LnM(η2‐C8H14) ( 6a—c ). The dinuclear cluster [Fe2(CO)6][CF‐CF‐C(t‐Bu)‐PH(OMe)] ( 8 ) was obtained from 3b and NiCl2·6H2O in methanol. The structures of 3a—d , 5a—c , 7b , and 8 have been elucidated by X‐ray crystal structure determinations.  相似文献   

17.
We report a new strategy to prepare a composite catalyst for highly efficient electrochemical CO2 reduction reaction (CO2RR). The composite catalyst is made by anchoring Au nanoparticles on Cu nanowires via 4,4′‐bipyridine (bipy). The Au‐bipy‐Cu composite catalyzes the CO2RR in 0.1 m KHCO3 with a total Faradaic efficiency (FE) reaching 90.6 % at ?0.9 V to provide C‐products, among which CH3CHO (25 % FE) dominates the liquid product (HCOO?, CH3CHO, and CH3COO?) distribution (75 %). The enhanced CO2RR catalysis demonstrated by Au‐bipy‐Cu originates from its synergistic Au (CO2 to CO) and Cu (CO to C‐products) catalysis which is further promoted by bipy. The Au‐bipy‐Cu composite represents a new catalyst system for effective CO2RR conversion to C‐products.  相似文献   

18.
Continued efforts are made for the utilization of CO2 as a C1 feedstock for regeneration of valuable chemicals and fuels. Mechanistic study of molecular (electro‐/photo‐)catalysts disclosed that initial step for CO2 activation involves either nucleophilic insertion or direct reduction of CO2. In this study, nucleophilic activation of CO2 by complex [(NO)2Fe(μ‐MePyr)2Fe(NO)2]2? ( 2 , MePyr=3‐methylpyrazolate) results in the formation of CO2‐captured complex [(NO)2Fe(MePyrCO2)]? ( 2‐CO2 , MePyrCO2=3‐methyl‐pyrazole‐1‐carboxylate). Single‐crystal structure, spectroscopic, reactivity, and computational study unravels 2‐CO2 as a unique intermediate for reductive transformation of CO2 promoted by Ca2+. Moreover, sequential reaction of 2 with CO2, Ca(OTf)2, and KC8 established a synthetic cycle, 2 → 2‐CO2 → [(NO)2Fe(μ‐MePyr)2Fe(NO)2] ( 1 ) → 2 , for selective conversion of CO2 into oxalate. Presumably, characterization of the unprecedented intermediate 2‐CO2 may open an avenue for systematic evaluation of the effects of alternative Lewis acids on reduction of CO2.  相似文献   

19.
The reagent Me3Si(C6F5) was used for the preparation of a series of perfluorinated, pentafluorophenyl‐substituted 3,6‐dihydro‐2H‐1,4‐oxazines ( 2 – 8 ), which, otherwise, would be very difficult to synthesize. Multiple pentafluorophenylation occurred not only on the heterocyclic ring of the starting compound 1 (Scheme), but also in para position of the introduced C6F5 substituent(s) leading to compounds with one to three nonafluorobiphenyl (C12F9) substituents. While the tris(pentafluorophenyl)‐substituted compound 3 could be isolated as the sole product by stoichiometric control of the reagent, the higher‐substituted compounds 5 – 8 could only be obtained as mixtures. The structures of the oligo(perfluoroaryl) compounds were confirmed by 19F‐ and 13C‐NMR, MS, and/or X‐ray crystallography. DFT simulations of the 19F‐ and 13C‐NMR chemical shifts were performed at the B3LYP‐GIAO/6‐31++G(d,p) level for geometries optimized by the B3LYP/6‐31G(d) level, a technique that proved to be very useful to accomplish full NMR assignment of these complex products.  相似文献   

20.
The ability to capture, store, and use CO2 is important for remediating greenhouse‐gas emissions and combatting global warming. Herein, Au nanoparticles (Au‐NPs) are synthesized for effective electrochemical CO2 reduction and syngas production, using polyethylenimine (PEI) as a ligand molecule. The PEI‐assisted synthesis provides uniformly sized 3‐nm Au NPs, whereas larger irregularly shaped NPs are formed in the absence of PEI in the synthesis solution. The Au‐NPs synthesized with PEI (PEI?Au/C, average PEI Mw=2000) exhibit improved CO2 reduction activities compared to Au‐NPs formed in the absence of PEI (bare Au NPs/C). PEI?Au/C displays a 34 % higher activity toward CO2 reduction than bare Au NPs/C; for example, PEI?Au/C exhibits a CO partial current density (jCO) of 28.6 mA cm?2 at ?1.13 VRHE, while the value for bare Au NPs/C is 21.7 mA cm?2; the enhanced jCO is mainly due to the larger surface area of PEI?Au/C. Furthermore, the PEI?Au/C electrode exhibits stable performance over 64 h, with an hourly current degradation rate of 0.25 %. The developed PEI?Au/C is employed in a CO2‐reduction device coupled with an IrO2 water‐oxidation catalyst and a proton‐conducting perfluorinated membrane to form a PEI?Au/C|Nafion|IrO2 membrane‐electrode assembly. The device using PEI?Au/C as the CO2‐reduction catalyst exhibits a jCO of 4.47 mA/cm2 at 2.0 Vcell. Importantly, the resulted PEI?Au/C is appropriate for efficient syngas production with a CO ratio of around 30–50 %.  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