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.
The inclusion complexation of aromatic amines with cucurbit[6]uril (CB[6]) capped with alkali metal cations was studied spectrophotometrically. We showed that CB[6] capped with alkali metal cations forms a 1:1 inclusion complex with the aromatic amine guests (neutral organic molecules), independent of the length of guest molecules. The effects of salts on the inclusion constants of CB[6] in the presence of different alkali salts were examined and it was found that the inclusion constants increased in the order of alkali cation Cs+ < Na+ < K+, suggesting the interaction of amine guests with the capped alkali metal cation. Further, the structures of the inclusion complexation of aromatic amines with CB[6] were characterized by 1H NMR measurements. Based on the results, the inclusion abilities of CB[6] capped with alkali metal cations are discussed.  相似文献   

2.
Channel-based reaction-diffusion equations are solved analytically for two electron transfer (ET) models, where the fast inner-sphere coordinate leads to an ET reaction treated by Fermi's golden rule, and the slow solvent coordinate moves via diffusion. The analytic solution has let us derive an ET rate constant that modifies the Marcus-Jortner formula by adding a constant alpha which we call a dynamic correction factor. The dynamic correction factor measures the effect of solvent friction. When the relaxation of solvent dynamics is fast, the dynamic correction can be neglected and the ET rate constant reduces to the traditional Marcus-Jortner formula. If the solvent dynamic relaxation is slow, the dynamic correction can be very large and the ET rate can be reduced by orders of magnitude. Using a generalized Zusman-Sumi-Marcus model as a starting point, we introduce two variants, GZSM-A and GZSM-B, where in model A, only one quantum mode is considered for inner-sphere motion and in model B, a classical mode for inner-sphere motion is added. By comparing the two models with experimental data, it is shown that model B is better than model A. For the solvents that have a relaxation time ranging between 0 and 5 ps, our result agrees fairly well with experimental data; for the solvents that have a relaxation time ranging between 5 and 40 ps, our result deviates from the experimental values. After introducing an adjustable scaling index in the effective time correlation function of the reaction coordinate, good agreement is achieved between the experiment and the theory for model B for all of the solvents studied in this paper.  相似文献   

3.
4.
Radiobrominations of cetyl bromide with82Br in several organic media such as DMF, 18-crown-6, DMSO, pyridine and n-butanol were achieved and a comparison was made as to the labelling rates. The rate order was found as follows: DMF>18-crown-6>DMSO>pyridine } n-butanol which was attributed to increased anion /i.e., the nucleophilic bromide ion,82Br/ solvation. Labelling in pyridine or alcohol was found to be too slow to meet practical needs. A suitable crown ether such as 18-crown-6 was found to be favourable for rapid radiobromination. However, it was discovered unexpectedly that dimethylformamide /DMF/ was superior to 18-crown-6. We therefore recommend DMF as an alternative reaction medium for fast radiobromination considering that this solvent is inexpensive and readily available.  相似文献   

5.
Gerardo Palla 《Tetrahedron》1981,37(16):2917-2919
The influence of reaction medium on the relative reactivity of various carbon radicals towards 4-CH3-and 4-CN- quinolines has been investigated. A relevant solvent effect for α-amidoalkyl and α-oxyalkyl radicals has been demonstrated and a solvent-radical interaction is proposed to explain the phenomenon.  相似文献   

6.
Reactions of nitrophenols have been studied in the eutectic melt of 8-hydroxyquinoline–benzoic acid, where it reacted with 8-hydroxyquinoline. The reactions were also carried out in solution. The reaction products obtained from both the methods were characterized by FT-IR, differential scanning calorimetry, X-ray diffraction technique and microstructural investigations. The reaction products obtained from eutectic melt were analyzed for C, H, N. The results showed that reaction products obtained from both the methods are same. An attempt has been made to propose the overall mechanism of the reaction in the eutectic melt.  相似文献   

7.
Rates and thermodynamic data have been obtained for the reversible self-termination reaction: Involving aromatic 2-(4′dimethylaminophenyl)indandione-1,3-yl (I), 2-(4′diphenylaminophenyl)indandione-1,3-yl (II), and 2,6 di-tert-butyl-4-(β-phthalylvinyl)-phenoxyl (III) radicals in different solvents. The type of solvent does not tangibly affect the 2k1 of Radical(I), obviously due to a compensation effect. The log(2k1) versus solvent parameter ET(30) curves for the recombination of radicals (II) and (III) have been found to be V shaped, the minimum corresponding to chloroform. The intensive solvation of Radical (II) by chloroform converts the initially diffusion-controlled recombination of the radical into an activated reaction. The log (2k?1) of the dimer of Radical (I) has been found to be a linear function of the Kirkwood parameter (ε - 1)/(2ε + 1), the dissociation rate increasing with the dielectic constant of the solvent. The investigation revealed an isokinetic relationship for the decay of the dimer of Radical (I), an isokinetic temperature β = 408 K and isoequilibrium relationship for the reversible recombination of Radical (I) with β° = 651 K. For Radical (I) dimer decay In(2k?1) = const + 0.8 In K, where K is the equilibrium constant of this reversible reaction. The transition state of Radical (I) dimer dissociation reaction looks more like a pair of radicals than the initial dimer. The role of specific solvation in radical self-termination reactions is discussed.  相似文献   

8.
Suzuki coupling of the bifunctional substrate 1 using [Pd(2) (dba)(3) ]/PtBu(3) gives selectivity for C?Cl in nonpolar solvents but for C?OTf in polar solvents. The results of computational and experimental studies suggest that the catalytically active species in polar solvents under conditions employing coordinating additives is inconsistent with monoligated [Pd(PtBu(3) )]. Instead, the data are consistent with an anionic palladium complex as the active species.  相似文献   

9.
Porcine pancreatic lipase-catalysed transesterifications of 2,2,2-trifluoroethyl butyrate with racemic 2-octanol and 1-phenylethanol have been studied in different organic solvents. Solvent hydrophobicity (log P -1.1 to 3.3) has only a minor effect on the reaction rate. Independently of the solvent used as the reaction medium, both (R)-2-octyl and (R)-1-phenylethyl butyrates were obtained in high optical purity (ee greater than 90%). Candida cylindracea lipase is active only in the most hydrophobic solvents studied.  相似文献   

10.
Gold-catalysed A3-reactions proceed efficiently when conducted in 2,2,2-trifluoroethanol as solvent. The rates of these reactions are accelerated considerably when conducted in a microwave reactor.  相似文献   

11.
12.
Conclusions In the systems: dialkyl phosphite (dialkyl thiophosphite)-alkyl thiocyanate-alkaline agent the unambiguous transformations via the Michaelis-Becker reactions proceed under the conditions of ion-dipole reactions with involvement of the unsolvated dialkyl phosphite or dialkyl thiophosphite anions. Sodium methylate, ammonia, and the alkali metal hydroxides can be used to form the latter in the presence of either dimethylformamide or hexamethylphosphoric triamide.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 12, pp. 2837–2840, December, 1974.  相似文献   

13.
14.
15.
A microscopic theory of solvent reorganization energy in polar molecular solvents is developed. The theory represents the solvent response as a combination of the density and polarization fluctuations of the solvent given in terms of the density and polarization structure factors. A fully analytical formulation of the theory is provided for a solute of arbitrary shape with an arbitrary distribution of charge. A good agreement between the analytical procedure and the results of Monte Carlo simulations of model systems is achieved. The reorganization energy splits into the contributions from density fluctuations and polarization fluctuations. The polarization part is dominated by longitudinal polarization response. The density part is inversely proportional to temperature. The dependence of the solvent reorganization energy on the solvent dipole moment and refractive index is discussed.  相似文献   

16.
17.
Solvent effects on the intrinsic activity of, the ion-exchange rate of, and the overall activity of polymer-supported phosphonium salts under tri-phase conditions were studied. The intrinsic activity of the catalysts as well as of soluble phosphonium salts was dependent only slightly on organic solvents. The exchange rate of the chloride ion in the catalysts against the acetate ion dependent on the solvents when the degree of ring substitution was ≤ 16%. With the ca. 30% ring-substituted catalysts the rate increased and hardly depended on the solvents. The overall reactivity of the catalysts for the reaction of organic halides with NaCN was a function of substrate structure and organic solvents. For alkyl halides such as bromooctane the catalysts were more reactive in good solvents such as toluene or chlorobenzene than in poor solvents such as octane. For arylalkyl halides such as benzyl chloride the catalysts exhibited the opposite effect. In poor solvents the arylalkyl halides are absorbed preferentially into the catalysts.  相似文献   

18.
Rate coefficients for the reactions of S,S-dimethyl-, S-methyl-S-phenyl- and S,S-diphenyl-N-tosylsulphilimine with benzenethiol in various alcohols have been determined at different temperatures. Multiparametric equations, involving specific and non-specific solute-solvent interactions, have been used to interpret quantitatively the data. The results on the solvent effect and the activation parameters indicate that the reactions proceed through the rate-determining formation of a thiol-sulphilimine adduct which occurs most likely through a cyclic transition state involving one molecule of thiol, one of sulphilimine, and one of alcohol.  相似文献   

19.
There were many discussions in the literature describing the membrane formation mechanism for the phase inversion process such as liquid–liquid demixing or crystallization, but few references described the phenomena after the event of the phase inversion process. The purpose of this work is to illustrate the effect of the second phase inversion on membrane structures when the first phase inversion has occurred. Analysis showed the second phase inversion (crystallization or liquid–liquid demixing) may be preceded by the first phase inversion (liquid–liquid demixing only) during poly (ethylene-co-vinyl alcohol) (EVAL) membrane formation. Therefore, we can make membranes combined with macrovoids (the first phase inversion) and particulate morphology (the second phase inversion) from experiments in this work. As a result, the concept the membrane morphology only influenced by the liquid–liquid demixing is misleading and the second phase inversion must be considered as a possible and important mechanism.  相似文献   

20.
The effects of replacing H2O with D2O solvent upon the electrochemical kinetics of simple transition-metal redox couples containing aquo, ammine or ethylenediamine ligands have been investigated at mercury electrodes as a means of exploring the possible contribution of ligand-aqueous solvent interactions to the activation barrier to outer-sphere electron transfer. The general interpretation of solvent isotope effects upon electrode kinetics is discussed; it is concluded that double-layer corrected isotopic rate ratios (kH/kD)E determined at a constant electrode potential vs. an aqueous reference electrode, as well as those determined at the respective standard potentials in H2O and D2O (kSH/kSD), have particular significance since the solvent liquid-junction potential can be arranged to be essentially zero. For aquo redox couples, values of (kSH/kSD) were observed that are substantially greater than unity and appear to be at least partly due to a greater solvent-reorganization barrier in D2O arising from ligand-solvent hydrogen bonding. For ammine and ethylenediamine complexes values of (kH/kD)E substantially greater than, and smaller than, unity were observed upon the separate deuteration of the ligands and the surrounding solvent respectively. Comparison of isotope rate ratios for corresponding electrochemical and homogeneous outer-sphere reactions involving cationic ammine and aquo complexes yields values of (kH/kD) for the former processes that are typically markedly larger than those predicted by the Marcus model from the homogeneous rate ratios. These discrepancies appear to arise from differences in the solvent environments in the transition states for electrochemical and homogeneous reactions.  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