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.
D ‐(+)‐Camphor forms the enamine 2 with piperidine. Compound 2 adds HB(C6F5)2 at the enamine carbon atom C3 to form a Lewis acid/Lewis base adduct (exo‐/endo‐isomers of 3 ). Exposure of 3 to dihydrogen (2.5 bar, room temperature) leads to heterolytic splitting of H2 to form the H+/H? addition products ( 4 , two diastereoisomers) of the “invisible” frustrated Lewis pairs ( 5 , two diastereoisomers) that were apparently generated in situ by enamine hydroboration under equilibrium conditions.  相似文献   

2.
The reactivity of dicoordinated Sn0 compounds, stannylones, is probed using density functional theory (DFT)‐based reactivity indices and compared with the reactivity of dicoordinated SnII compounds, stannylenes. For the former compounds, the influence of different types of electron‐donating ligands, such as cyclic and acyclic carbenes, stannylenes and phosphines, on the reactivity of the central Sn atom is analyzed in detail. Sn0 compounds are found to be relatively soft systems with a high nucleophilicity, and the plots of the Fukui function f? for an electrophilic attack consistently predict the highest reactivity on the Sn atom. Next, complexes of dicoordinated Sn compounds with different Lewis acids of variable hardness are computed. In a first part, the double‐base character of stannylones is demonstrated in interactions with the hardest Lewis acid H+. Both the first and second proton affinities (PAs) are high and are well correlated with the atomic charge on the Sn atom, probing its local hardness. These observations are also in line with electrostatic potential plots that demonstrate that the tin atom in Sn0 compounds bears a higher negative charge in comparison to SnII compounds. Stannylones and stannylenes can be distinguished from each other by the partial charges at Sn and by various reactivity indices. It also becomes clear that there is a smooth transition between the two classes of compounds. We furthermore demonstrate both from DFT‐based reactivity indices and from energy decomposition analysis, combined with natural orbitals for chemical valence (EDA‐NOCV), that the monocomplexed stannylones are still nucleophilic and as reactive towards a second Lewis acid as towards the first one. The dominating interaction is a strong σ‐type interaction from the Sn atom towards the Lewis acid. The interaction energy is higher for complexes with the cation Ag+ than with the non‐charged electrophiles BH3, BF3, and AlCl3.  相似文献   

3.
The silica-supported iron Lewis acid [(η5-C5H5)Fe(CO)2(THF)]+ [BF4]+ (1), was found to catalyze the formation of an enol ester, cyclopropane, or aziridine from a diazo compound and an aldehyde, olefin, or imine, respectively. Unlike the non silica-supported iron Lewis acid, the catalyst is recoverable and reusable.  相似文献   

4.
Electrochemical characteristics of MF-4SK/PAni nanocomposite membranes prepared at different times of chemical polymerization of aniline are studied. Electroosmotic permeability and conductivity of membranes in solutions of acids and sodium chloride are determined. It is revealed that the conductivity of nanocomposites in the proton form at 30-day synthesis is approximately 3 times as low as that of the base membrane and composite membrane formed at 5-h synthesis. The transport number of water slightly depends on the structural type of membrane and changes from 3.3 to 2 mol H2O/mol H+ with an increase in the concentration of HCl solution from 0.1 to 3 M. The ratio of transport number to the water content rises about twofold in composites as compared to the initial membrane. It is shown that water is transferred with proton as hydronium structures [H5O2]+ and [H9O4]+ by the migration mechanism whose contribution to the total proton transfer in composite membranes increases.  相似文献   

5.
The reactivity of the hydrolysis product of hexaphenylcarbodiphosphorane, PPh3CHP(O)Ph2, towards different soft Lewis acids, such as CuI and Ag[BF4] are reported. While CuI exclusively binds at the ylidic carbon atom, reaction of the silver cation in CH2Cl2 leads to proton abstraction from the solvent to give the cation [PPh3CH2P(O)Ph2]+. Surprisingly, Ag+ replaces the methyl group of [PPh3CHMeP(O)Ph2]+ to produce a dimeric complex, in which Ag+ is coordinated to C and O forming an eight membered ring. The compounds were characterized by spectroscopic methods and X‐ray diffraction.  相似文献   

6.
《印度化学会志》2021,98(8):100101
Although a helium atom prefers to stay at the centre of a fullerene (C60) cage and a proton binds with one of the carbon atoms from inside, DFT(MN15)/cc-pVTZ and DLPNO-MP2/def2-TZVP calculations show that the helium atom and the proton in HeH+ prefer to stay away from the centre of the cage, weakening the He–H+ covalent bond considerably. Both the helium atom and the proton exhibit noncovalent interactions with the carbon atoms of two pentagons at the opposite ends of the fullerene cage. Our calculations also show that a linear arrangement of H3+ (inside C60), pointing towards the centres of two pentagons opposite to each other, with the proton breaking away from H2, is energetically more favored over the equilateral triangle geometry of free H3+.  相似文献   

7.
A comb-like copolymer consisting of a poly(vinylidene fluoride-co-chlorotrifluoroethylene) backbone and poly(hydroxy ethyl acrylate) side chains, i.e. P(VDF-co-CTFE)-g-PHEA, was synthesized through atom transfer radical polymerization (ATRP) using CTFE units as a macroinitiator. Successful synthesis and a microphase-separated structure of the copolymer were confirmed by proton nuclear magnetic resonance (1H NMR), FT-IR spectroscopy, and transmission electron microscopy (TEM). This comb-like polymer was crosslinked with 4,5-imidazole dicarboxylic acid (IDA) via the esterification of the –OH groups of PHEA and the –COOH groups of IDA. Upon doping with phosphoric acid (H3PO4) to form imidazole–H3PO4 complexes, the proton conductivity of the membranes continuously increased with increasing H3PO4 content. A maximum proton conductivity of 0.015 S/cm was achieved at 120 °C under anhydrous conditions. In addition, these P(VDF-co-CTFE)-g-PHEA/IDA/H3PO4 membranes exhibited good mechanical properties (765 MPa of Young's modulus), and high thermal stability up to 250 °C, as determined by a universal testing machine (UTM) and thermal gravimetric analysis (TGA), respectively.  相似文献   

8.
We have found that hydrogen uranyl phosphate tetrahydrate HUO2PO4·4H2O has a high proton conductivity. The ac conductivity was 0.4 ohm?1 m?1 at 290°K measured parallel to the faces of sintered disks of the compound. The activation energy was found to be 31 ± 3 kJ mole?1. The values of conductivity were between 3 and 10 times lower when measured perpendicular to the disk faces due to preferred orientation of the plate-like crystals. Both the powder and sintered disks are stable in air and insoluble in phosphoric acid solution of pH 2.5. Experiments are described which enable possible grain boundary contributions to the conductivity to be determined in such hydrates. The extrinsic grain boundary contribution to the conductivity was found to be small from experiments in which the pH in a solution cell was varied. The abnormally high bulk H+ conductivity thus inferred is attributed primarily to the high concentration of H+, which exists as H3O+ in the interlamellar hydrogen-bonded network. A Grotthus-type mechanism of conduction is proposed which involves intermolecular transfer steps (hopping) and intramolecular transfer steps, in comparable numbers, the former facilitated by the high concentration of H3O+ ions in the structure, and the latter most likely facilitated by the high concentration of H-bond vacancies.  相似文献   

9.
The kinetics of reaction of the [W3PdS4H3(dmpe)3(CO)]+ hydride cluster ( 1 +) with HCl has been measured in dichloromethane, and a second‐order dependence with respect to the acid is found for the initial step. In the presence of added BF4? the second‐order dependence is maintained, but there is a deceleration that becomes more evident as the acid concentration increases. DFT calculations indicate that these results can be rationalized on the basis of the mechanism previously proposed for the same reaction of the closely related [W3S4H3(dmpe)3]+ cluster, which involves parallel first‐ and second‐order pathways in which the coordinated hydride interacts with one and two acid molecules, and ion pairing to BF4? hinders formation of dihydrogen bonded adducts able to evolve to the products of proton transfer. Additional DFT calculations are reported to understand the behavior of the cluster in neat acetonitrile and acetonitrile–water mixtures. The interaction of the HCl molecule with CH3CN is stronger than the W? H???HCl dihydrogen bond and so the reaction pathways operating in dichloromethane become inefficient, in agreement with the lack of reaction between 1 + and HCl in neat acetonitrile. However, the attacking species in acetonitrile–water mixtures is the solvated proton, and DFT calculations indicate that the reaction can then go through pathways involving solvent attack to the W centers, while still maintaining the coordinated hydride, which is made possible by the capability of the cluster to undergo structural changes in its core.  相似文献   

10.
Although the existence of the NF4+ cation has been known for 51 years, and its formation mechanism from NF3 , F2 , and a strong Lewis acid in the presence of an activation energy source had been studied extensively, the mechanism had not been established. Experimental evidence had shown that the first step involves the generation of F atoms from F2 , and also that the NF3+ cation is a key intermediate. However, it was not possible to establish whether the second step involved the reaction of a F atom with either NF3 or the Lewis acid (LA). To distinguish between these two alternatives, a computational study of the NF4 , SbF6 , AsF6 , and BF4 radicals was carried out. Whereas the heats of reaction are small and similar for the NF4 and LAF radicals, at the reaction temperatures, only the LAF radicals possess sufficient thermal stability to be viable species. Most importantly, the ability of the LAF radicals to oxidize NF3 to NF3+ demonstrates that they are extraordinary oxidizers. This extraordinary enhancement of the oxidizing power of fluorine with strong Lewis acids had previously not been fully recognized.  相似文献   

11.
Hydrated (undecomposed) form of heteropolyacid H3PMo12O40/SiO2 exhibits a higher activity in the formic acid decomposition than the corresponding dehydrated sample. The formic acid decomposition takes place on strong Br?nsted acid sites of the heteropolyacid.Ab initio SCF MO LCAO method was used for the calculation of the electronic state of two surface complexes of HCOOH molecule (S1 and S2) coordinated to a proton H+. The S1 complex is formed by proton addition to the carbonyl oxygen, whereas the S2 complex is formed proton addition to the oxygen atom of the C−O−H fragment of HCOOH. The selective weakening of the C−O bond and localization of the positive charge on the (O=C−H) fragment in the protonated complex S2 are favorable for the decomposition of formic acid to CO and H2O.  相似文献   

12.
The catalytic system Ni(COD)2/BF3·OEt2 is highly active in the addition polymerization of nor-bornene (NB). Its activity, which is up to 1930 (kg NB) (mol Ni)−1 h−1, is higher than the activity of the other known nickel complex catalysts. Another advantage of this system over the latter is that it contains a smaller proportion of a Lewis acid (5 molar parts or below) and no conventional stabilizing organoelement ligands. The activity of this system in NB polymerization has been investigated by Fourier-transform IR spectroscopy. According to EPR data, NB polymerization is accompanied by the formation of low-spin complexes of trivalent nickel, which result from the oxidative addition of the monomer to univalent nickel complexes. A metallacyclic mechanism involving Ni(I) and Ni(III) complexes is suggested for NB polymerization.  相似文献   

13.
In this study, a novel bidentate ligand containing oxime, hydrazone, and indole moieties and its BF2+-bridged transition metal complexes [Ni(II), Cu(II), and Co(II)] were synthesized and their cytotoxic activities against prostate and breast cancer cells were investigated. The vic-dioxime ligand bearing indole–hydrazone side groups was synthesized by reacting antiglyoximehydrazine (GH2) with 3-methoxy indole. The ligand forms mononuclear complexes with a metal-to-ligand ratio of 1:2 with M = Co(II)(H2O)2, Ni(II), and Cu(II). These metal complexes were then reacted with BF3(C2H5)2O to obtain BF2+-bridged transition metal complexes. The Co(II) complex of the ligand is proposed to be octahedral with water molecules as axial ligands, whereas the Ni(II) and Cu(II) complexes are proposed to be square planar. Spectral studies showed that the ligand bonded to the metal ion in a neutral bidentate fashion through the azomethine nitrogen atom and the imine oxime group. Structural assignments are supported by a combination of 1H nuclear magnetic resonance (NMR), 13C NMR, Fourier-transform infrared, LC/MS, elemental analyses, and magnetic susceptibility testing. For determining the cytotoxic effects of the novel anticancer products, cancer cells were cultured. The antiproliferative effects were determined using the MCF-7 breast cancer and PC-3 prostate cancer cell lines. The antiproliferative effects of the products were analyzed and their apoptotic or necrotic effects were determined with the Hoechst/propidium iodide double staining method in both cancer cell lines. Paclitaxel was used as the positive control (1 μm ). The results indicated that the newly synthesized compounds are effective on both cell lines between concentrations of 5 and 40 μm and show their effects by apoptotic mechanisms. Besides, these products were found to be more effective on the MCF-7 cell line. The cytotoxic efficiency of the newly synthesized products was more than that of paclitaxel (depending on concentration), which is a chemotherapeutic agent used in cancer therapy.  相似文献   

14.
Boron trifluoride (BF3) is a highly corrosive gas widely used in industry. Confining BF3 in porous materials ensures safe and convenient handling and prevents its degradation. Hence, it is highly desired to develop porous materials with high adsorption capacity, high stability, and resistance to BF3 corrosion. Herein, we designed and synthesized a Lewis basic single-crystalline hydrogen-bond crosslinked organic framework (HCOF-50) for BF3 storage and its application in catalysis. Specifically, we introduced self-complementary ortho-alkoxy-benzamide hydrogen-bonding moieties to direct the formation of highly organized hydrogen-bonded networks, which were subsequently photo-crosslinked to generate HCOFs. The HCOF-50 features Lewis basic thioether linkages and electron-rich pore surfaces for BF3 uptake. As a result, HCOF-50 shows a record-high 14.2 mmol/g BF3 uptake capacity. The BF3 uptake in HCOF-50 is reversible, leading to the slow release of BF3. We leveraged this property to reduce the undesirable chain transfer and termination in the cationic polymerization of vinyl ethers. Polymers with higher molecular weights and lower polydispersity were generated compared to those synthesized using BF3 ⋅ Et2O. The elucidation of the structure–property relationship, as provided by the single-crystal X-ray structures, combined with the high BF3 uptake capacity and controlled sorption, highlights the molecular understanding of framework-guest interactions in addressing contemporary challenges.  相似文献   

15.
Controlling the reactivity of transition metals using secondary, σ‐accepting ligands is an active area of investigation that is impacting molecular catalysis. Herein we describe the phosphine gold complexes [(o‐Ph2P(C6H4)Acr)AuCl]+ ([ 3 ]+; Acr=9‐N‐methylacridinium) and [(o‐Ph2P(C6H4)Xan)AuCl]+ ([ 4 ]+; Xan=9‐xanthylium) where the electrophilic carbenium moiety is juxtaposed with the metal atom. While only weak interactions occur between the gold atom and the carbenium moiety of these complexes, the more Lewis acidic complex [ 4 ]+ readily reacts with chloride to afford a trivalent phosphine gold dichloride derivative ( 7 ) in which the metal atom is covalently bound to the former carbocationic center. This anion‐induced AuI/AuIII oxidation is accompanied by a conversion of the Lewis acidic carbocationic center in [ 4 ]+ into an X‐type ligand in 7 . We conclude that the carbenium moiety of this complex acts as a latent Z‐type ligand poised to increase the Lewis acidity of the gold center, a notion supported by the carbophilic reactivity of these complexes.  相似文献   

16.
Controlling the reactivity of transition metals using secondary, σ‐accepting ligands is an active area of investigation that is impacting molecular catalysis. Herein we describe the phosphine gold complexes [(o‐Ph2P(C6H4)Acr)AuCl]+ ([ 3 ]+; Acr=9‐N‐methylacridinium) and [(o‐Ph2P(C6H4)Xan)AuCl]+ ([ 4 ]+; Xan=9‐xanthylium) where the electrophilic carbenium moiety is juxtaposed with the metal atom. While only weak interactions occur between the gold atom and the carbenium moiety of these complexes, the more Lewis acidic complex [ 4 ]+ readily reacts with chloride to afford a trivalent phosphine gold dichloride derivative ( 7 ) in which the metal atom is covalently bound to the former carbocationic center. This anion‐induced AuI/AuIII oxidation is accompanied by a conversion of the Lewis acidic carbocationic center in [ 4 ]+ into an X‐type ligand in 7 . We conclude that the carbenium moiety of this complex acts as a latent Z‐type ligand poised to increase the Lewis acidity of the gold center, a notion supported by the carbophilic reactivity of these complexes.  相似文献   

17.
The formation of a frustrated Lewis pair consisting of sodium hydride (Na+H?) and a framework‐bound hydroxy proton O(H+) is reported upon H2 treatment of zeolite NaY loaded with Pt nanoparticles (Ptx/NaY). Frustrated Lewis pair formation was confirmed using in situ neutron diffraction and spectroscopic measurements. The activity of the intrazeolite NaH as a size‐selective catalyst was verified by the efficient esterification of acetaldehyde (a small aldehyde) to form the corresponding ester ethyl acetate, whereas esterification of the larger molecule benzaldehyde was unsuccessful. The frustrated Lewis pair (consisting of Na+H? and O(H+)) generated within zeolite NaY may be a useful catalyst for various catalytic reactions which require both H? and H+ ions, such as catalytic hydrogenation or dehydrogenation of organic compounds and activation of small molecules.  相似文献   

18.
The polymerization of ethylene on a chromic oxide catalyst with and without a solvent has been studied. It was found that the active catalyst surface is formed exclusively as a result of its interaction with ethylene. This interaction is accompanied by the formation of products which poison the surface of the catalyst when they are sorbed on it in the absence of a solvent. A catalyst which contains no Cr+6 atoms as a result of reduction by alcohol is inactive. On the other hand, a catalyst which contains only Cr+6 atoms becomes active only after it has been partially reduced. The most probable product of this reduction is trivalent chromium atoms. The results obtained have given grounds for the assumption that the active complex contains Cr+6 and Cr+3 atoms. A possible mechanism of the reaction is discussed. Owing to the oxidative action of CrO3 on the ethylene molecules, part of the Cr+6 is reduced to Cr+3, and the trivalent chromium becomes alkylated. The monomer molecule is added at the Cr+3—C bond thus formed. A strong Lewis acid, CrO3, lowers the electron density on the Cr+3 atom. This increases the strength of the Cr+3—C bond and the ability of the Cr+3 atom to coordinate with the monomer molecule. The monomer molecule enters the chain at the moment when the strength of the Cr?3—C bond is weakened due to coordination of this molecule with the Cr+3 atom.  相似文献   

19.
It is shown by quantum chemical simulation (MP2/aug-cc-pVTZ) that the energy of addition of H+, H·, and H? decreases in the order ethylene, CO, and N2. The energies of additions of CF4, dimethyl ether, and BF3 to the ions and radicals formed were calculated. Unlike the CH3CH2 ? ion, the HCO? ion can add exothermically not one, but two BF3 molecules.  相似文献   

20.
The basic study on the determination of tetrafluoroborate ion (BF4) by ion chromatography, and total boron by conversion of boric acid to BF4 followed by ion chromatography of BF4 has been carried out. The results of thermodynamic calculations for the system of boric acid (H3BO3)-F-H+ showed that the mole fraction of BF4 was higher than 99% at pH lower than 3.5 and 4.5 when the total free fluoride concentration (2[H2F2] + 2[HF2] + [HF] + [F]) was as high as 0.1 and 1.0 M, respectively. The fraction of BF4 increased with increasing total free fluoride concentration. BF4 fraction values were higher than 99% at pH 0.75 and at total free fluoride concentration of 0.05 M or higher. BF4 was hardly formed at pH > 7 even when the total free fluoride concentration was as high as 1.0 M. According to the experimental results, the fraction of BF4 at pH 0.7-0.8 was 51.2, 95.6 and 96.7% when the total fluoride concentration (2[H2F2] + 2[HF2] + [HF] + [F] + 3[BF3OH] + 4[BF4]) was 0.2, 1.0 and 3.3 M, respectively. The formation reaction of BF4 from boric acid reached an equilibrium state within 20 min regardless of reaction temperature, in the range of 20-50 °C, when the total boron and total fluoride concentrations were 66.7 mM and 1.0 M, respectively. Although BF4 was formed only under acidic conditions, BF4, once formed, was very stable under alkaline conditions at least for several hours. We have concluded that BF4 could be analyzed by ion chromatography using sodium hydroxide solution as an eluent because BF4 was stable under chromatographic conditions. BF4 solution prepared from boric acid could be used as a standard solution in the ion chromatographic analysis of BF4 instead of the sodium tetrafluoroborate (NaBF4) reagent available commercially, if a discrepancy of about 4-5% was allowed.  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