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 593 毫秒
1.
The rate constants and activation parameters of the reactions of the carbocation resulting from 6-ethoxy-1,2,2,4-tetramethyl-1,2-dihydroquinoline photolysis with methanol (k 1) and the methoxide ion (k 2) have been measured by flash photolysis in binary mixtures of methanol with inert solvents (nonpolar pentane and polar acetonitrile) in wide composition ranges. The changes in the activation parameters for k 1 at different solvent compositions show that the increase in the rate constant in the pentane mixtures is mainly deter-mined by the increase in the preexponential factor. The decrease in k 1 in the acetonitrile mixtures is deter-mined by the decrease in the methanol concentration and by the increase in the activation energy. The different roles of the methoxide ion in the reaction are demonstrated. They depend on the nature of the inert solvent in the mixture. The results of this study are considered in terms of methanol clustering in pentane and acetonitrile, the different solubilities of 6-ethoxy-1,2,2,4-tetramethyl-1,2-dihydroquinoline in the components of the binary mixtures, and the difference in distribution and solvation between the carbocation and the methoxide ion in the mixtures.  相似文献   

2.
The dependence of the quantum yield and the decay rate constant for the carbocation generated in the photolysis of 6-ethoxy-1,2,2,4-tetramethyl-1,2-dihydroquinoline (6-EtO-DHQ) in methanol on the solvent composition was studied in the mixtures of methanol with isopropanol, acetonitrile, and pentane by pulse photolysis. The addition of these solvents decreases the yield of the carbocation and differently affects the kinetic parameters of its decay. The carbocation decay in the mixtures MeOH-i-PrOH and MeOH-C5H12 is described by the pseudo-first order equation (k 1), with the dependence of k 1 having a maximum at 50 vol % of MeOH in the MeOH-i-PrOH mixtures, and k 1 increasing with a decrease in the MeOH concentration in the MeOH-C5H12 mixtures. In the MeOH-MeCN mixtures, the value of k 1 decreases with a decrease in the MeOH concentration, and, at the concentration of MeOH lower than 50 vol %, the contribution of the second-order reaction (k 2) is observed. The activation energies and preexponential factors were determined in the MeOH-C5H12 mixtures of different compositions, and it was shown that E act practically did not depend on the solvent composition and were close to E act for other carbocations obtained in MeOH. The increase in k 1 with a decrease in the MeOH concentration is caused by an increase in the preexponential factor. The results were discussed on the basis of the reaction mechanism involving the competing reactions of the carbocation combination with two nucleophilic particles, the MeOH molecule and the MeO anion. The composition of the mixture and the nature of the inert solvent affect strongly the course of these reactions. Published in Russian in Kinetika i Kataliz, 2009, Vol. 50, No. 3, pp. 411–416. The article was translated by the authors.  相似文献   

3.
The rate constants of the reactions of diphenyl carbonyl oxide Ph2COO with a number of alcohols and water in acetonitrile, benzene, andn-decane solutions (295 K) were measured by flash photolysis. The rate constants vary over a range from 400 (triphenylmethanol in a MeCN solution) to 2.5 × 105 l mol–1 s–1 (adamantanol in a benzene solution). -Methoxydiphenylmethyl hydroperoxide is the reaction product of Ph2COO and MeOH. The absence of a kinetic isotope effect and the dependence of the logarithms of the rate constants on the first ionization potentials of alcohols are indicative of the formation of a C–O bond at the rate-limiting step of the reaction.  相似文献   

4.
The reaction of the azide ion with the carbocation generated in the photolysis of 1,2,2,4,6-pentamethyl-1,2-dihydroquinoline in methanol was studied by pulse (conventional and laser) and steady-state photolysis techniques. The adduct of the azide ion was characterized by 1H NMR spectrum. Experimental results were interpreted taking into account a competition between the addition of methanol and azide ion to the carbocation. The rate constants for the reaction of the azide ion with the carbocation (k Az) were measured at 2—48 °C in a wide range of [N3 ]0 concentrations from 2·10–7 to 0.1 mol L–1 at different ionic strengths () of the solution. The resulting k Az values are more than an order of magnitude lower than those for diffusional-controlled reactions and vary from 3.2·108 ( = 0) to 4.5·106 L mol–1 s–1 ( = 0.8 mol L–1) in the presence of NaClO4 (18 °C). The activation energy of addition of the azide ion to the carbocation is 21 kJ mol–1, which is by 12 kJ mol–1 lower than the activation energy of the reaction of the carbocation with methanol. The features of the reaction under study are discussed from the viewpoint of the structures of carbocations generated in the photolysis of dihydroquinolines.  相似文献   

5.
Benzyl chloride and benzyl acetate were photolyzed in 30% methanol–water mixtures (V/V) at 0°C. The photolysis produces benzyl carbocations that react with nucleophiles. The reaction products were analyzed by gas chromatography or liquid chromatography. From the amounts of products the relative values of rate constants of reactions of benzyl carbocation with nucleophiles N and water k(N)/k(H2O) were calculated. Benzyl carbocation reacts with I?, Br?, Cl?, and Ac? ions with approximately diffusion-controlled rate. A value of 2.4 × 107 dm3 mol?1 s?1 for the rate constant k(H2O) and a lifetime of 0.7 ns were estimated for benzyl carbocation in the aqueous solution.  相似文献   

6.
The photolysis of 2,2,4,6-tetramethyl-1,2-dihydroquinoline in binary MeOH-C5H12 and MeOH-MeCN mixtures was studied in dependence on the solvent composition by steady state and pulse photolysis. The photoinduced proton transfer from the N-H bond to the C(3) carbon atom of the heterocycle does not occur up to the methanol concentration of 0.25 (1 vol %) and 2.5 mol l−1 (10 vol %) in the MeOH-C5H12 and MeOH-MeCN mixtures, respectively. The trends in the increase in the relative quantum yield of the proton transfer reaction (Q mix/Q MeOH) and the decay rate constant for 2,2,4,6-tetramethyl-2,3-dihydroquinoline, the product of the proton transfer reaction, in dependence on the composition differ significantly for these binary mixtures. The results are interpreted in terms of peculiarities in the aggregation of methanol and the distribution of the DHQ molecules and transient species in the mixtures.  相似文献   

7.
The efficiency of formation and the decay kinetics of carbocations formed under the photolysis of 1,2,2,4-tetramethyl-1,2-dihydroquinoline in methanol and in a porous glass filled with methanol or dried in air or in vacuo were studied by the laser flash photolysis techniques. In MeOH, the carbocations recombine via the second-order law in the reaction with the MeO anion formed in an equimolar amount and decay via the first-order law in the reaction with the solvent with rate constants of 3·108 L mol−1 s− 1 and 1.4·103 s−1, respectively. When the solution is placed into the porous glass, no recombination of the carbocations with MeO is observed, and the reaction with the solvent is somewhat inhibited (rate constant 8·102 s−1). More than tenfold inhibition of the reaction of the carbocations with methanol is observed on going to a monolayer of MeOH on the surface. The main route of carbocation decay in the porous glass dried in vacuo is the geminate recombination with the SiO groups. The corresponding kinetics is described in terms of the model of freely diffusing reactants. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2239–2243, October, 2005.  相似文献   

8.
The acid-catalysed hydrolysis of potassium ethyl malonate in presence of 0.05N-HCl in water and in dioxane-water as well as acetone-water mixtures containing up to about 90% by weight of the organic solvent was studied over the temperature range 30–55°C. The reaction rate decreases continuously with decreasing water content of the solvent mixtures until it reaches a minimum at about 80% dioxane and increases again. The isocomposition activation energy shows only slight changes with solvent composition. The kinetics of the reaction was investigated from the viewpoints of solvent composition, water concentration and dielectric constant. The radii and thermodynamic parameters of the activated complex were calculated and discussed. The conclusions drawn show that the reaction is better treated as an ion-dipole rather than an ion-ion interaction.With 1 Figure  相似文献   

9.
The kinetics of the photolysis of substituted 1,2-dihydroquinolines (DHQ) in micellar solutions was studied by steady-state and flash photolysis. The photolysis mechanism depends dramatically on the location of DHQ molecules in micelles, which is governed by the surfactant nature. In micellar solutions of the anionic surfactant sodium dodecyl sulfate (SDS), where the DHQ molecules are located in the Stern layer, the intermediate species decay kinetics follows a first-order law. When DHQ is in neutral form (pH 4–12), the rate constant of the intermediate carbocation decay increases from 25 to 198 s?1 with an increasing concentration of DHQ in micelles. The positive micellar catalysis is caused by the acceleration of the final product formation with the DHQ molecule via proton abstraction from the intermediate cation. The formation of several types of intermediate species—carbocations in the aqueous phase and aminyl radicals in micelles—is observed in micellar solutions of the cationic surfactant cetyltrimethylammonium bromide (CTAB) due to the preferential location of DHQ molecules in the micellar core. The carbocation decays via a pseudofirst-order reaction with a rate constant close to that in the aqueous solution. The lifetime of the DHQ aminyl radicals in the micellar solutions is longer by several orders of magnitude than the lifetime observed for homogeneous solutions of hydrocarbons and alcohols.  相似文献   

10.
The flash photolysis of diphenyldiazomethane in acetonitrile, benzene, and n-decane solutions saturated with air resulted in the formation of diphenyl carbonyl oxide Ph2COO which decayed in combination reactions. In the presence of organic sulfides, the transfer of the terminal oxygen atom of Ph2COO to the sulfur atom was observed. The kinetics of this reaction was studied. The absolute rate constants (k 6, dm3 mol–1 s–1) of the reactions of Ph2COO with sulfides at 295 K (acetonitrile as a solvent) varied from 4.1 × 102 (Me2S) to 8.1 × 104 (Ph2S). The solvent effect on the reaction kinetics and product composition was studied. The mechanism of the process was discussed.  相似文献   

11.
Summary The kinetics of the thermal decomposition of thetris-oxalactocobaltate(III) complex has been investigated in the presence of EtOH and (CH2OH)2 spectrophotometrically in the 50–70 ± 0.1 ° C range. The rate of the reaction decreases upon the addition of either of the alcohols to the reaction medium, whereas the activation energy increases. The thermodynamic parameters were calculated and are discussed in terms of the solvation effects. The influence of the dielectric constants of the solvent mixtures on the rate has been studied. A free radical mechanism is proposed and discussed.  相似文献   

12.
The photolysis of o-azidobenzoic acid and its potassium salt in water, ethanol, tetrahydrofuran (THF), and ethanol–water and THF–water mixtures was studied by IR and UV spectroscopy and thin-layer chromatography. It was found that the photolysis of o-azidobenzoic acid and potassium o-azidobenzoate in aqueous solutions mainly resulted in 2,1-benzisoxazolone, the intramolecular cyclization product. The process of azepine formation in ethanol or an ethanol–water mixture depends on the nucleophilic nature of ethanol; in this case, water is practically of no importance. The presence of THF in an aqueous solution shifted the singlet nitrene–dehydroazepine equilibrium toward dehydroazepine, which reacts with water to give azepines.  相似文献   

13.
Summary The kinetics of the solvolysis of the ions [Co(NH3)5Cl]2+ and [Co(CN)5Cl]3– have been investigated in a range of water-rich mixtures of water with n-PrOH at various temperatures. The variation of the enthalpies and entropies of activation with solvent composition show extrema in composition ranges where the physical properties of the mixtures, influenced by changes in solvent structure, also show extrema. From the application of a free energy cycle to the process of the initial state going to the transition state, it is concluded for the solvolyses of both complexes that the species of CoIII in the transition state is more stable in water + n-PrOH mixtures than the ionic complexes of CoIII in the initial state. The results are compared with those for the solvolysis of these and related complexes in mixtures of water with other alcohols.  相似文献   

14.
The dependence of the rate constants for the oxidation of [Co(en)2amS]2+ (amS=–SCH2CH2NH2 2– by IO4– on solvent composition has been established for aqueous mixtures of MeOH, i-PrOH, t-BuOH, urea (U), ethylene carbonate (EC), acetonitrile (AN), dimethyl formamide (DMF) and acetone (Me2CO). Addition of aprotic cosolvents leads to marked decrease in the reaction rate in comparison to monohydroxylic alcohols. Alcohols inhibit the reaction rate in order MeOH < i-PrOH < t-BuOH. The change in solvation on going from the initial to the transition state are discussed in terms of the Gibbs transfer function, Gt0.  相似文献   

15.
Rate constants and derived thermodynamic activation parameters are reported for solvolysis of trans-[Co(3Mepy)4Cl2]+ and [Co(CN)5Cl]3– ions in water-rich mixtures of water with ethanol at various temperatures and are analyzed by initial- and transition-state contributions. The variation of enthalpies and entropies of activation with solvent composition show extrema in composition ranges where the physical properties of the mixtures, influenced by changes in solvent structures, also show extrema. From the application of a free-energy cycle to the process of the initial state going to the transition state, it is concluded for the solvolysis of both complexes that the Co(III) species in the transition state is more stable in water + ethanol mixtures than in the initial state.  相似文献   

16.
The kinetics of oxidation of a series of monoatomic alcohols (methanol, 2-propanol, 2-butanol, 2-methyl-1-propanol, 2-chloroethanol, 1,3-dichloro-2-propanol, benzyl alcohol), hydroxyacetic acid, and 1,3-butandiol (ROH) by dimethyldioxirane (1) was studied. The reaction kinetics obeys the second order equationw=k[ROH][1]. The rate constants were measured in the range of 7–50†C, and the activation parameters were found. To describe the reaction rate constants as a function of the ROH structure, the two-parametric Taft equation was used, which takes into account both the polar and resonance substituent effects. Alcohol oxidation produces the corresponding carbonyl compounds,viz., ketones from secondary alcohols and aldehydes from primary alcohols, in yields of at least 80%. The results were explained by the competition of the molecular (oxenoid) and radical mechanisms. The introduction of electron-withdrawing substituents into the alcohol molecule increases the contribution of the radical channel of the reaction. Published inIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1338–1343, August, 2000.  相似文献   

17.
The mechanism of the photoinduced Markovnikov addition of water and methanol to the double bond of 2,2,4-trimethyl-8-methoxy-1,2-dihydroquinoline was studied by flash photolysis over a wide range of acid and alkali concentrations. In neutral MeOH solutions, the successive formation of two active intermediates was observed. These intermediates have absorption bands at max = 420 and 500 nm, respectively, and lifetimes of a few milliseconds. The rate constants of decay for these species (k 1 and k 2) were measured over the temperature range from 10 to 45°C, and the activation energies E 1 (3.7 ± 0.4) and E 2 (6.9 ± 0.7) kcal/mol were determined in MeOH. The addition of an acid accelerates the conversion of the former intermediate into the latter. The rate constant for the reaction of the former intermediate with the acid is (2.0 ± 0.4) × 108 l mol–1 s–1. The latter intermediate was proposed to be a corresponding carbocation. In H2O at pH < 9, the formation of the carbocation (max = 500 nm) was observed on a millisecond timescale, with the lifetime being an order of magnitude higher than that in MeOH. In alkaline media, only the former intermediate (max = 420 nm) was observed with a lifetime of 150 or 400 ms in MeOH or H2O, respectively, at [NaOH] > 2 × 10–3 mol/l. The quantum yields of the product and the quantum yields of fluorescence were measured as functions of acidity. Possible mechanisms of the reaction are discussed.  相似文献   

18.
1-Norbornyloxychlorocarbene (1-NorOCCl), 1-bicyclo[2.2.2]octyloxychlorocarbene (1-BcoOCCl), and 1-adamantyloxychlorocarbene (1-AdOCCl) were generated in dichloroethane (DCE) by photolysis of the appropriate diazirines. The exclusive product in each case was the bridgehead alkyl chloride formed by fragmentation of the carbene to [R(+) OC Cl(-)] ion pairs, loss of CO, and cation-anion collapse. In mixtures of methanol and DCE, each carbene gave three products: RCl, ROH, and ROMe. RCl and ROMe resulted from competition between ion pair collapse and methanol capture of the cation. ROH resulted from methanol capture of the carbene (before fragmentation), followed by eventual methanolysis and hydrolysis of ROCH(Cl)OMe. The ratios of carbene capture to carbene fragmentation fell in the order 1-NorOCCl > BcoOCCl > 1-AdOCCl; 1-Nor(+) was the least stable cation and the slowest to form by fragmentation, so that this carbene was the most readily captured. This trend was accentuated in methanol-pentane mixtures, where ionic fragmentation was further slowed in the less polar solvent. Laser flash photolysis with either UV or time-resolved infrared (TRIR) monitoring permitted the determination of the absolute rate constants for fragmentations of the carbenes in DCE at 25 degrees C. The rate constants (s(-1)) were: 1-NorOCCl (3.3 x 10(4)), 1-BcoOCCl (1.5 x 10(5)), and 1-AdOCCl (5.9 x 10(5)). The rate constants decreased in the order of increasing strain in the resulting bridgehead carbocation, but the range of rate constants was compressed to a factor of only approximately 18. This constrasts with the factor of 10(10) by which the acetolysis of 1-AdOTs at 70 degrees C exceeded that of 1-NorOTs. The fragmentation of 1-NorOCCl to the ion pair was 3 x 10(15) times faster than the acetolysis of 1-NorOTs. The activation energies were measured as 9.0 kcal/mol (log A = 11.2 s(-1)) for the fragmentation of 1-NorOCCl and 4.4 kcal/mol (log A = 8.44 s(-1)) for that of 1-BcoOCCl both in DCE. B3LYP/6-31G computed activation energies in simulated DCE were 14.6 and 2.7 kcal/mol, respectively.  相似文献   

19.
Pulsed Corona Discharge-Induced Reactions of Acetophenone in Water   总被引:2,自引:0,他引:2  
The reactions of acetophenone in water by pulsed corona discharges have been investigated to provide fundamental information concerning the reactions of acetophenone in water. Experimental results indicated that photolysis of acetophenone did not involve a hydroxyl radical mechanism and the majority flux of hydroxyl radicals originated from the dissociation of gas-phase oxygen in the plasma channels. The rate constants for photolysis and pyrolysis were determined to be 1.5×10–7 M-s–1, 2.2×10–4 s–1, respectively. The rate constant for the oxidative reactions was measured as 1.2×10–7 M-s–1. Results from this study support the proposal that acetophenone degradation reaction proceed through the oxidative reaction pathway, where molecular oxygen accelerates acetophenone degradation, photolysis, and pyrolysis pathways.  相似文献   

20.
Free, unsolvated phenylium ions formed by the spontaneous β decay of [1,4-3H2]benzene have been allowed to react with gaseous alcohols (ROH: R = Et, CF3CH2, Pr, or i-Pr; partial pressure: 3–56 Torr) and alkyl chlorides (R′Cl: R′ = Pr, i-Pr, or Bu; partial pressure: 20–450 Torr), in the presence of a thermal radical scavenger (O2: 4 Torr). Phenylium ion confirms its considerable site selectivity, demonstrated by the distinct preference toward the n-centre of the substrate (46–100%), although significant insertion into the alkyl group of alcohols is observed as well. Phenylium ion displays significant positional selectivity even between different n-type sites in a bidentate molecule such as CF3CH2OH. An affinity F < O < Cl trend is observed, which indicates a direct relationship between the polarizability of the n-centre of the molecule and its orienting properties toward phenylium ion. The stability features of the ionic intermediates from addition of phenylium ion with ROH or R′Cl have been evaluated, as well as their fragmentation and isomerization mechanisms. The behaviour of phenylium ion toward the selected substrates in the gas phase is discussed and compared with previous mechanistic hypotheses from related nuclear-decay studies.  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