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.
Abstract

The reaction between tetraphenylphosphonium chloride and hydroxide or deuteroxide anions was studied kinetically in a series of dimethylsulphoxide-water mixtures at several temperatures. The rate is first-order in the phosphonium cation and second-order in the hydroxide or deuteroxide anions. The reaction shows a dramatic increase in rate, up to about 1010 times, as the DMSO content is increased. The rate enhancement is attributed to a considerable drop in activation energy affected not only through an increased desolvation of reactant anions, but also through an increase in solvation of the transition state, brought about by gradual addition of DMSO. The kinetic solvent deuterium isotope effect in 60% DMSO-40% D2O is strongly dependent on temperature. The rate constant in the latter solvent mixture is represented by k i = 11.9 e ?12700/RT l 2 mole?2 sec?1 as compared to k i = 19.0 e ?22500/RT l 2 mole?2 sec?1 in the corresponding 60% DMSO-H2O mixture. The thermodynamic parameters of activation show strong dependence on solvent composition and are related to structural changes and solvation power of the reaction medium.  相似文献   

2.
Rate constants for the reaction of O(3P) atoms with C3H4, C3H6 and NO(M = N2O) have been measured over the temperature range 300–392°K using a modulation-phase shift technique. The Arrhenius expressions obtained are:C2H4, k2 = 3.37 × 109 exp[?(1270 ± 200)/RT]liter mole?1 sec?1,C3H6, k2 = 2.08 × 109 exp[?(0 ± 300)/RT]liter mole?1 sec?1,NO(M = N2O), k1 = 9.6 × 109 exp[(900 ± 200/RT]liter2 mole?2 sec?1.These temperature dependencies of k2 are in good agreement with recent flash photolysis-resonance flourescence measurements, although lower than previous literature values.  相似文献   

3.
By allowing dimethyl peroxide (10?4M) to decompose in the presence of nitric oxide (4.5 × 10?5M), nitrogen dioxide (6.5 × 10?5M) and carbon tetrafluoride (500 Torr), it has been shown that the ratio k2/k2′ = 2.03 ± 0.47: CH3O + NO → CH3ONO (reaction 2) and CH3O + NO2 → CH3ONO2 (reaction 2′). Deviations from this value in this and previous work is ascribed to the pressure dependence of both these reactions and heterogeneity in reaction (2). In contrast no heterogeneous effects were found for reaction (2′) making it an ideal reference reaction for studying other reactions of the methoxy radical. We conclude that the ratio k2/k2′ is independent of temperature and from k1 = 1010.2±0.4M?1 sec?1 we calculate that k2′ = 109.9±0.4M?1 sec?1. Both k2 and k2′ are pressure dependent but have reached their limiting high-pressure values in the presence of 500 Torr of carbon tetrafluoride. Preliminary results show that k4 = 10.9.0±0.6 10?4.5±1.1M?1 sec?1 (Θ = 2.303RT kcal mole?1) and by k4 = 108.6±0.6 10?2.4±1.1M?1 sec?1: CH3O + O2 → CH2O + HO2 (reaction 4) and CH3O + t-BuH → CH3OH + (t-Bu) (reaction 4′).  相似文献   

4.
The rate of oxidation of Ge(II) chloride by large excess of ClO2? ions in HCl, NaCl and Na2SO4 mixed solutions was polarographically observed at various H2O+ and Cl? ion concentrations. The observed rate constant, kobs, is expressed by ko=Kobs/(ClO3?)={k1,(H+)+k2K1(Cl?)2+ K3K2(SO42?)} (H+)/{(H+)1+K1(Cl-)2 +K2(SO42?)} for the following reaction processes, The values were obtained aa k1=1.5410-3liter2 mole2? sec-1, k2=5.00×10-2liter2 mole2? sec-2 and k2=4.30×10-3liter2 mole2? sec-2, K1=1.80× 10-2, K2= 2.43×10-2 mole liter-1 at constant ionic strength I=0.50 M at 30°C.  相似文献   

5.
The rate constant for the combination of trichloromethyl radicals in the gas phase has been measured by applying the rotating sector technique to the gas phase carbon tetrachloride–cyclohexane photochemical system. A temperature-independent rate constant, k5, of 3.9 ± 1.0 × 1012 cc mole?1 sec?1 was found. Arrhenius parameters for the reaction were found to be given by the expression log k4 = 11.79 – (10,700/2.3 RT).  相似文献   

6.
The gas phase, nitric oxide catalyzed positional isomerization of 3-methylene-1,5,5-trimethylcyclohexene (MTC) into 1,3,5,5-tetramethyl-1,3-cyclohexadiene (TECD) has been studied for temperatures ranging between 296° and 425°C. The major reaction was first order with respect to nitric oxide and to MTC. The major side product, mesitylene, usually amounted to less than 10% of the TECD isomer formed. Only at high temperatures and large conversions has up to 20% been observed. Conditioned pyrex or quartz vessels coated with KCl have been used. The nitric oxide catalyzed isomerization is apparently a homogeneous process, as demonstrated by the insensitivity of the observed rate constants towards a 15-fold increase in the surface to volume ratio of the reaction vessels. However, a residual, presumably heterogeneous, thermal isomerization of the starting material could not be eliminated. Good mass balances were obtained for both NO and hydrocarbons. After correcting for the thermally induced conversion the observed rate constants for the nitric oxide catalyzed isomerization yield log k1 (1 mole?1 sec?1) = (10.7 ± 0.2) – (37.3 ± 0.9)/θ where θ is 2.303 × 10?3 RT (kcal mole?1). Plotting log k1 versus the ratio of the starting materials (MTC/NO)0 it was found that for temperatures ≥ 365°C the rate constants were systematically too high. Using extrapolated values for the higher temperature range yields the more reliable corrected Arrhenius equation log k = 8.6 – 31.7/θ. The reaction mechanism is outlined and the implications with respect to the stabilization energy generated in the MTC? radical intermediate and the activation energy of the backreaction MTC? + HNO are discussed. Using for the activation energy E?1 of the backreaction (R? + HNO) a literature value of 9.2 ± 0.9 kcal mole?1 reported for the cyclohexadiene? 1,3? system, this yields 23.4 ± 2 kcal mole?1 for the stabilization energy in the methylenecyclohexenyl radical, which is to be compared with the corresponding values for the allyl (10.2 ± 1.4), methallyl (12.6 ± 1) pentadienyl (15.4 ± 1) and cyclohexadienyl (24.6 ± 0.7) radicals. The pre-exponential factor agrees well with the value of (8.4 ± 0.2) reported by Shaw and co-workers for the similar reaction of NO with 1,3-cyclohexadiene. It is noteworthy that HNO, acting as sole hydrogen donor in the system, is surprisingly stable under the reaction conditions used. Nitrous oxide, HCN, H2O and N2 are observed in the product mixture of experiments carried out to high conversions at higher temperatures.  相似文献   

7.
4-Chloro-1-butene, 5-chloro-1-pentene, and 6-chloro-1-hexene have been shown to decompose, in a static system, mainly to hydrogen chloride and the corresponding alkadienes. In packed and unpacked clean Pyrex vessels the reactions were significantly heterogeneous. However, in a vessel seasoned with allyl bromide these reactions were homogeneous, unimolecular, and follow a first-order law. The working temperature range was 389.6–480.0°C and with a pressure range of 53–221 Torr. The rate constants for the homogeneous reactions were expressed by the following Arrhenius equations: 4-chloro-1-butene: logk(sec?1) = (13.79 ± 0.17) – (223.8 ± 2.1)kJ/mole/2.303RT; 5-chloro-1-pentene: logk(sec?1) = (14.25 ± 1.20) – (238.4 ± 12.7)kJ/mole/2.303RT; and 6-chloro-1-hexene: logk(sec?1) = (12.38 ± 0.22) – (209.6 ± 2.9)kJ/mole/2.303RT. The olefinic double bond has been found to participate in the rate of dehydrohalogenation of 4-chloro-1-butene. The insulation of the CH2?CH in chlorobutene by one or two methylene chains to the reaction center does not indicate neighboring group participation. The three-membered conformation is the most favored structure for anchimeric assistance of the double bond in gas phase pyrolysis of alkenyl chlorides. The heterolytic nature of these eliminations is also supported by the present work.  相似文献   

8.
It is shown by pulse radiolysis that in aqueous solutions of hydrazine containing oxygen the radical N2H3 reduces oxygen to O2 at pH > 7 (k 3·109 dm3· mole–1·sec–1), while this reaction does not occur for the protonated form N2H4 + at pH < 7 (k, 5·106 dm3·mole–1·sec–1). The rate constants for the disappearance of O2 have been determined in the pH range from 4 to 12. Rate constants have been calculated for the reaction of O with N2H4 [k=(1.6 ±0.2)·109 dm3·mole–1·sec–1] and of O3with N2H4 [k=(1.2 ±0.2)·106 dm3· mole–1·sec–1].Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 2, pp. 341–345, February, 1991.  相似文献   

9.
2,2,4-Trimethyl-3-on-1-pentyl methacrylate (TMPM) was first synthesized from the condensation reaction of 2,2,4-trimethyl-1-pentanol-3-on with methacrylic acid. Second, the polymerization of TMPM and the copolymerization of TMPM with styrene (St) were carried out in benzene at 60°C, using 2,2′-azobisisobutyronitrile (AIBN) as an initiator. As the result of kinetic investigation, the rate of polymerization (Rp) could be expressed by: Rp = k[AIBN]0.5 [TMPM]1.0. Kinetic constants of polymerization of TMPM were obtained as follows: kp/k = 0.27 dm3/2 mole?1/2 sec?1/2, 2fkd = 1.23 × 10?5 sec?1, f = 0.73, Cm = 2.6 × 10?5, Cs = 1.1 × 10?5. From the results the reactivity of TMPM was found to be larger than that of methyl methacrylate. The overall activation energy was calculated to be 110 kJ mole?1. The following monomer reactivity ratios and Q, e values were obtained: TMPM(M1) ? St(M2): r1 = 1.50, r2 = 0.14, Q1 = 2.63, E1 = 0.45.  相似文献   

10.
Abstract

The kinetics and stability constants of l-tyrosine complexation with copper(II), cobalt(II) and nickel(II) have been studied in aqueous solution at 25° and ionic strength 0.1 M. The reactions are of the type M(HL)(3-n)+ n-1 + HL- ? M(HL)(2-n)+n(kn, forward rate constant; k-n, reverse rate constant); where M=Cu, Co or Ni, HL? refers to the anionic form of the ligand in which the hydroxyl group is protonated, and n=1 or 2. The stability constants (Kn=kn/k-n) of the mono and bis complexes of Cu2+, Co2+ and Ni2+ with l-tyrosine, determined by potentiometric pH titration are: Cu2+, log K1=7.90 ± 0.02, log K2=7.27 ± 0.03; Co2+, log K1=4.05 ± 0.02, log K2=3.78 ± 0.04; Ni2+, log K1=5.14 ± 0.02, log K2=4.41 ± 0.01. Kinetic measurements were made using the temperature-jump relaxation technique. The rate constants are: Cu2+, k1=(1.1 ± 0.1) × 109 M ?1 sec?1, k-1=(14 ± 3) sec?1, k2=(3.1 ± 0.6) × 108 M ?1 sec?1, k?2=(16 ± 4) sec?1; Co2+, k1=(1.3 ± 0.2) × 106 M ?1 sec?1, k-1=(1.1 ± 0.2) × 102 sec?1, k2=(1.5 ± 0.2) × 106 M ?1 sec?1, k-2=(2.5 ± 0.6) × 102 sec?1; Ni2+, k1=(1.4 ± 0.2) × 104 M ?1 sec?1, k-1=(0.10 ± 0.02) sec?1, k2=(2.4 ± 0.3) × 104 M ?1 sec?1, k-2=(0.94 ± 0.17) sec?1. It is concluded that l-tyrosine substitution reactions are normal. The presence of the phenyl hydroxyl group in l-tyrosine has no primary detectable influence on the forward rate constant, while its influence on the reverse rate constant is partially attributed to substituent effects on the basicity of the amine terminus.  相似文献   

11.
The mechanism and kinetics of the γ-ray-initiated postpolymerization of octadecyl methacrylate and acrylate in lamellar crystals were investigated by a simple model. This model assumes that the initiation points are distributed as in a checkerboard and that polymerization probability of the monomer molecules decreases conically around each initiation point. The two-dimensional polymerization can be characterized in this cone model by two parameters, a and r; a represents the polymerizability of the monomer for a given condition, and r depends on the number of initiation points per unit area. G values for the initiation reaction of octadecyl methacrylate and acrylate were estimated as 0.8 and 1.6, respectively. The two-dimensional postpolymerization of long-chain compounds proceeds in two stages. The rate of polymerization is very high and zero order with respect to monomer concentration in the first stage. It is lower and obeys first-order kinetics in the second stage. The rate constants of the zero-and first-order polymerizations were kp0 = 1.73 molecule sec-1 and kp1 = 0.93 sec?1, respectively, for octadecyl acrylate at 20°C.  相似文献   

12.
The kinetics of the reaction of O + CH3OCH3 were investigated using fast-flow apparatus equipped with ESR and mass-spectrometric detection. The concentration of O(3P) atoms to CH3OCH3 was varied over an unusually large range. The rate constant for reaction was found to be k = (5.0 ± 1.0) × 1012 exp [(?2850 ± 200/RT)] cm3 mole?1 sec?1. The reaction O + CH3OH was studied using ESR detection. Based on an assumed stoichiometry of two oxygen atoms consumed per molecule of CH3OH which reacts, we obtain a value of k = (1.70 ± 0.66) × 1012 exp [(?2,280 ± 200/RT)] cm3 mole?1 sec?1 for the reaction The results obtained in this study are compared with the results from other workers on these reactions. The observation of essentially equal activation energies in these two reactions is indicative of approximately equal C? H bond strengths in CH3OCH3 and CH3OH. This is in agreement with recent measurements of these bond energies.  相似文献   

13.
Reactions of 2,4,6-tri-t-butylphenoxyl (TBP) with cumene hydroperoxide (ROOH), cumylperoxyl radicals (RO2), and molecular oxygen in benzene solution have been investigated kinetically by the ESR method. The rate constant of the reaction TBP + ROOH has been estimated in the temperature range 27°-75°C: log10(k?7/M?1sec?1) = (7.1 ± 0.4) - (10.9 ± 0.6 kcal mole?1)/θ The ratio of the rate constants of reactions TBPH + RO2 products has been determined from the experimental dependence of the rate constant of reaction TBP with ROOH on [TBPH]0/[TBP]0. Putting k7 = 4.0 × 103M?1sec?1, we obtain k8 = (2.0 ± 0.2) × 108M?1sec?1 at 30°C. The reaction of TBP with O2 obeys the kinetic law ?d[TBP]/dt = k′[O2][TBP]2. This is in accordance with scheme TBP + O2 ← TBP ?O2 [I]; TBP ?O2 + TBP · products, log10 (k′/M?2sec?1) = (?14.5 ± 0.9) + (27.2 ± 1.4)/θ at 66°?78°C, where ° = 2.303RT.  相似文献   

14.
The kinetics of the reaction of cis-(NO)2 with solid oxygen to form iso-N2O4 have been studied between 13 and 29 K. The overall reaction is pseudo first order in cis-(NO)2, and solid oxygen serves both as a reactant and the matrix. The pseudo-first-order rate constants are calculated to be k(14N) = 4.25 × 10?2 exp(-103/RT), and k(15N) = 3.00 × 10?2 exp(-105/RT) sec?1, based on temperature measurements from a thermocouple junction which may be at most three degrees lower than the actual reacting film. Most significantly, however, 14k/15k = 1.55 at~13 K. The condensed phase reaction has been compared to that observed in the gas phase, and the extremely small pre-exponential factors and large isotope effects have been discussed in terms of tunneling corrections and orientational constraints. It is suggested that the form of the crystal plays an integral role in the observed process.  相似文献   

15.
An absolute value of kr of ethyl radicals at 860 ± 17°K of 4.5 × 109 M?1·sec?1 was determined under VLPP conditions, where the value of kr/kr should be about 1/2. Thus kr(M?1·sec?1) ~ 1010 at 860°K. An error of as much as a factor of 2 in kr would be surprising, but possible. The value of 1010M?1·sec?1 seems to be a factor of from 2 to 5 too high to be compatible with extensive data on the reverse reaction and the accepted thermochemistry. Changes in the heat of formation and entropy of the ethyl radical can change the situation somewhat, but even these changes when applied to the work of Hiatt and Benson [3] indicate that ethyl combination should be ~ 109.3 M?1·sec?1. More work is necessary if a better value is desired.  相似文献   

16.
The Co(NH3)5OH23+ ion reacts with malonate to form Co(NH3)5O2CCH2CO2H2+ or Co(NH3)5O2CCH2CO2+, depending on the pH of the reaction solution. The kinetics of this anation reaction have been studied as a function of [H+] for the acidity range 1.5 ≤ pH ≤ 6.0 in the temperature range of 60 to 80°C, the [total malonate] ≤ 0.5 M, and the ionic strength 1.0M. The anation by malonic acid follows second-order kinetics, the rate constant being 8.0 × 10?5 M?1·sec?1 at 70°C, and the anations by bimalonate (Q1, k1) and malonate ion (Q2, k2) are consistent with an Id mechanism. Typical values at 70°C for the ion pair formation constants are Q1 = 1.3, Q2 = 5.4M?1; and for the interchange rate constants k1 = 5.3 × 10?4; k2 = 7.3 × 10?4 sec?1. The activation parameters for the various rate constants are reported and the results discussed with reference to previously reported data for similar systems.  相似文献   

17.
The thermal isomerization of the title compounds was studied in the vapor phase. Over the temperature range from 445.1 to 477.5°K, 1,4-dimethylbicyclo[2.2.0]hexane underwent a homogeneous unimolecular reaction to 2,5-dimethyl-1,5-hexadiene, the rate constants being represented by the equation: k = 1.86 × 1011 exp (?31000 ± 1800/RT) sec?1. Over the temperature range from 630.0 to 662.2°K, 1,4-dimethylbicyclo[2.1.1]-hexane also underwent a unimolecular isomerization to the same product, the rate constants being given by the equation: k = 8.91 × 1014 exp (?56000 ± 900/RT) sec?1. The pyrolysis of 1,4-dimethylbicyclo[2.1.0]pentane gave 1,3-dimethylcyclopentene-1 and 2,4-dimethyl-1,4-pentadiene in the ratio of 9:1. The former reaction was influenced by surface effects but the latter was not. The rate constants for the formation of 2,4-dimethyl-1,4-pentadiene fitted the equation: k = 1.66 × 1017 exp (?57400 ± 3100/RT) sec?1. The effect of the two methyl groups at the bridgehead positions in these molecules in influencing the rate of decomposition is discussed in terms of the non-bonded repulsive forces between the substituents.  相似文献   

18.
The kinetics of the gas-phase dehydrogenation of cyclopentane to cyclopentene is found to be consistent with a slow attack by an I atom (step 4, text) on cyclopentane in the range 282–382°C. The measured rate constants fit the Arrhenius equation, log k4 = 11.95 ± 0.08 – (24.9 ± 0.23)/θ 1 mole?1 sec?1, where θ = 2.303 R T in kcal/mole. This leads to a value of ΔH = 24.3 ± 1 kcal/mole and a bond dissociation energy DH = 94.9 ± 1 kcal/mole. The latter value is identical with DH0(i-Pr-H) = 95 ± 1 kcal/mole and signifies that cyclopentane and the cyclopentyl radical have the same strain energy. Arrhenius parameters are deduced for all six steps in the reaction mechanism. Surface reactions are shown to be unimportant. Cyclopentyl iodide is an unstable intermediate in the reaction and the rate constant for its bimolecular formation from HI + cyclopentene is found to be log k6 = 8.40 ± 0.29 - (26.9 ± 0.8)/θ 1 mole?1 sec?1. Together with the equilibrium constant, this yields for the unimolecular elimination of HI from cyclopentyl iodide, the rate constant, log k5 = 13.3 ± 0.3 – (42.8 ± 1.2)/θ sec?1.  相似文献   

19.
The kinetics of the thermally and radiation initiated chain reaction between trichloroethylene and cyclopentane to produce 1,1-dichlorovinylcyclopentane and hydrogen chloride have been investigated in the temperature range 250–360°C at high pressure in the gas phase. The rate governing step in the chain is (k3 = 3.3 × 109 exp ?(4800/RT) cc mole?1 sec ?1). The rate of the unimolecular decomposition of trichloroethylene is 1.4 × 1014 exp ?(61,200/RT) sec?1.  相似文献   

20.
Azoethane was irradiated in the presence of carbon monoxide in the temperature range of 238 to 378 K. Kinetic parameters for the addition of ethyl radicals to carbon monoxide and for the decomposition of propionyl radicals were determined. The rate constants were found to be log k(cm3 mol?1 sec?1) = 11.19 - 4.8/θ and log k(sec?1) = 12.77 - 14.4/θ, respectively. Estimated thermochemical properties of the propionyl radical are ΔHf0 = -10.6 ± 1.0 kcal mol?1, S0 = 77.3 ± 1.0 cal K?1 mol?1, and D(C2H5CO? H) = 87.4 kcal mol?1.  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