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.
The reaction mechanisms for oxidation of CH3CCl2 and CCl3CH2 radicals, formed in the atmospheric degradation of CH3CCl3 have been elucidated. The primary oxidation products from these radicals are CH3CClO and CCl3CHO, respectively. Absolute rate constants for the reaction of hydroxyl radicals with CH3CCl3 have been measured in 1 atm of Argon at 359, 376, and 402 K using pulse radiolysis combined with UV kinetic spectroscopy giving ??(OH + CH3CCl3) = (5.4 ± 3) 10?12 exp(?3570 ± 890/RT) cm3 molecule?1 s?1. A value of this rate constant of 1.3 × 10?14 cm3 molecule?1 s?1 at 298 K was calculated using this Arrhenius expression. A relative rate technique was utilized to provide rate data for the OH + CH3 CCl3 reaction as well as the reaction of OH with the primary oxidation products. Values of the relative rate constants at 298 K are: ??(OH + CH3CCl3) = (1.09 ± 0.35) × 10?14, ??(OH + CH3CClO) = (0.91 ± 0.32) × 10?14, ??(OH + CCl3CHO) = (178 ± 31) × 10?14, ??(OH + CCl2O) < 0.1 × 10?14; all in units of cm3 molecule?1 s?1. The effect of chlorine substitution on the reactivity of organic compounds towards OH radicals is discussed.  相似文献   

2.
The reactions of tert-butoxyl radicals with amines, leading to the formation of α-aminoalkyl radicals, and the reactions of these with the electron acceptor methyl viologen have been examined using laser flash photolysis techniques. For example, the radicals CH3?HNEt2 and HOCH2?H N(CH2CH2OH)2 react with methyl viologen with rate constants equal to (1.3 ± 0.1) × 109 and (2.1 ± 0.4) × 109M?1 · s?1, respectively, in wet acetonitrile at 300 K.  相似文献   

3.
Absolute rate constants have been measured for the reactions of trichloromethylperoxyl radicals with cyclohexane, cyclohexene, and hexamethylbenzene. The CCl3O2 radicals were produced by pulse radiolysis of air-saturated CCl4 solutions containing various amounts of the hydrocarbons. The rate constants were determined by competition with the one-electron oxidation of metalloporphyrins, using the rate of formation of the metalloporphyrin radical cation absorption to monitor the reaction by kinetic spectrophotometry. The rate constants for hydrogen abstraction from cyclohexane, cyclohexene, and hexamethylbenzene were found to be 1 × 103, 1.0 × 105, and 7.5 × 104 M?1 s?1, respectively.  相似文献   

4.
Trichloromethylperoxyl radicals were produced by pulse radiolysis of air saturated solutions containing CCl4. The rate constants for the reaction of CCl3O2 radicals with zinc tetraphenylporphyrin (ZnTPP) were determined in various solvents. They were found to vary between 3 × 107 and 3 × 109 M?1 s?1. The changes in rate constants result from complexation of ZnTPP with the different solvents, but did not correspond to changes in redox potential of ZnTPP. The rate constants were found to depend on the strength of the axial complexation, indicating an inner sphere mechanism whereby the radical binds to the metal prior to electron transfer.  相似文献   

5.
Rate constants for the tri-n-butyltin radical ( Sn · ) induced decomposition of a number of peroxides have been measured in benzene at 10°C. The values range from ~100 M?1 sec?1 for di-t-butyl peroxide to 2.6 × 107 M?1 sec?1 for di-t-butyl diperoxyisophthalate. The majority of the peroxides, including diethyl peroxide, diacetyl peroxide, and t-butyl peracetate, have rate constants of ~105 M?1 sec?1. It is shown that di-n-alkyl disulfides are ten times as reactive toward Sn · as di-n-alkyl peroxides, although the exothermicities of these reactions are ~15 and ~39 kcal/mole, respectively. The enhanced reactivity of the disulfides is attributed to the easier formation of an intermediate or transition state with 9 electrons around sulfur, compared with an analogous species with 9 electrons around oxygen. The following bond strengths (kcal/mole) have been estimated: D[ Sn ? OR] = 77; D[ Sn ? H] = 82; D[ Sn ? SR] = 83; and D[ Sn ? OC(O)R] = 86, where R = alkyl. Rate constants for reaction of Sn · with some benzyl esters have also been measured. It has been found that t-butoxy radicals can add to benzene and abstract hydrogen from benzene at ambient temperatures.  相似文献   

6.
Rate constants have been measured by pulse radiolysis for the reactions of the NO3 radical with five cyclic ethers and a series of alcohols. Rate constants ranged from 3.5 × 104 M×1 s×1 for deuterated methanol to 1.1 × 107 M?1 s?1 for tetrahydrofuran. The rate constants for the reactions of NO3 with the alcohols 1-propanol to 1-heptanol were found to be linearly dependent on the number of CH3 groups with a group reactivity of 6.4 × 105 M?1 s?1.  相似文献   

7.
Rate constants for a series of alcohols, ethers, and esters toward the sulfate radical (SO4?) have been directly determined using a laser photolysis set‐up in which the radical was produced by the photodissociation of peroxodisulfate anions. The sulfate radical concentration was monitored by following its optical absorption by means of time resolved spectroscopy techniques. At room temperature the following rate constants were derived: methanol ((1.6 ± 0.2) × 107 M?1 s?1); ethanol ((7.8 ± 1.2) × 107 M?1 s?1); tert‐butanol ((8.9 ± 0.3) × 105 M?1 s?1); diethyl ether ((1.8 ± 0.1) × 108 M?1 s?1); MTBE ((3.13 ± 0.02) × 107 M?1 s?1); tetrahydrofuran (THF) ((2.3 ± 0.2) × 108 M?1 s?1); hydrated formaldehyde ((1.4 ± 0.2) × 107 M?1 s?1); hydrated glyoxal ((2.4 ± 0.2) × 107 M?1 s?1); dimethyl malonate (CH3OC(O)CH2C(O)OCH3) ((1.28 ± 0.02) × 106 M?1 s?1); and dimethyl succinate (CH3OC(O)CH2CH2C(O)OCH3) ((1.37 ± 0.08) × 106 M?1 s?1) where the errors represent 2σ. For the two latter species, we also measured the temperature dependence of the corresponding rate constants. A correlation of these kinetics with the bond dissociation energy is also presented and discussed. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 539–547, 2001  相似文献   

8.
The kinetics and absolute rate constants for the free-radical chain reaction of tri-n-butyltin hydride with di-t-butyl disulfide have been measured in cyclohexane at 30°. The rate controlling step for chain propagation involves the cleavage of the disulfide bond by an attacking tributyltin radical. The rate constant for this bimolecular homolytic substitution at sulfur is ~8 × 104 Mole?1 sec?1. Chain termination involves the self-reaction of two tributyltin radicals. The rate constants for attack of tributyltin radicals on some other disulfides and on elemental sulfur have also been measured. The results are compared with literature data for homolytic substitutions on these compounds by a variety of radicals which have their unpaired electron centered on carbon.  相似文献   

9.
Rate constants for one-electron oxidation by the methylperoxyl radicals (CH3O2, HOCH2O2, ?O2CCH2O2, and CCI3O2) in aqueous solutions have been measured by pulse radiolysis and found to be in the range of 3 × 105 to 6 × 108 M?1 s?1 for compounds with redox potentials between 0.6 and 0.1 V. Substitution on the methylperoxyl radical with OH or CO2? has only a minor effect on the rate of oxidation but substitution with three chlorines increases the rate constants by two orders of magnitude. The redox potential of the CH3O2 radical is estimated to be 0.6–0.7 V.  相似文献   

10.
Formation of cation radicals by pulse radiolysis of metalloporphyrins and chlorophyll a in 1,2-dichloroethane is reported Demetalation of the metalloporphyrin by radiolytically produced HC1 is also observed. Rate constants for demetalation of ZnTPP and Chl a are 1 1 × 108 and ≈ 3 × 108 M?1 S?1. Oxidation of Chl a by ZnTPP+ has a rate constant of ≈ 4× 109 M?1S?1.  相似文献   

11.
The rate constant for the reaction has been determined by means of vacuum ultraviolet flash photolysis and time-resolved kinetic spectroscopic observations of the 1504-Å absorption band of CH3. The measurements made using three different sources of methyl radicals (azomethane, dimethylmercury, and ketene-hydrogen) were in accord and yielded a value for the rate constant of k1 = (9.53 ± 1.17) × 10?11 cc molec?1 sec?1. A detailed error analysis is presented. The f-value for the 1504-Å band of CH3 is determined to be (2.5 ± 0.7) × 10?2.  相似文献   

12.
The rate constants k1 for the reaction of CF3CF2CF2CF2CF2CHF2 with OH radicals were determined by using both absolute and relative rate methods. The absolute rate constants were measured at 250–430 K using the flash photolysis–laser‐induced fluorescence (FP‐LIF) technique and the laser photolysis–laser‐induced fluorescence (LP‐LIF) technique to monitor the OH radical concentration. The relative rate constants were measured at 253–328 K in an 11.5‐dm3 reaction chamber with either CHF2Cl or CH2FCF3 as a reference compound. OH radicals were produced by UV photolysis of an O3–H2O–He mixture at an initial pressure of 200 Torr. Ozone was continuously introduced into the reaction chamber during the UV irradiation. The k1 (298 K) values determined by the absolute method were (1.69 ± 0.07) × 10?15 cm3 molecule?1 s?1 (FP‐LIF method) and (1.72 ± 0.07) × 10?15 cm3 molecule?1 s?1 (LP‐LIF method), whereas the K1 (298 K) values determined by the relative method were (1.87 ± 0.11) × 10?15 cm3 molecule?1 s?1 (CHF2Cl reference) and (2.12 ± 0.11) × 10?15 cm3 molecule?1 s?1 (CH2FCF3 reference). These data are in agreement with each other within the estimated experimental uncertainties. The Arrhenius rate constant determined from the kinetic data was K1 = (4.71 ± 0.94) × 10?13 exp[?(1630 ± 80)/T] cm3 molecule?1 s?1. Using kinetic data for the reaction of tropospheric CH3CCl3 with OH radicals [k1 (272 K) = 6.0 × 10?15 cm3 molecule?1 s?1, tropospheric lifetime of CH3CCl3 = 6.0 years], we estimated the tropospheric lifetime of CF3CF2CF2CF2CF2CHF2 through reaction with OH radicals to be 31 years. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 36: 26–33, 2004  相似文献   

13.
Rate constants have been determined for the reactions of SO4? with a series of alcohols, including hydrated formaldehyde. The SO4? radical was produced by the laser-flash photolysis of persulfate, S2O82?. Rate constants for the reactions of SO4? with alcohols range from 1.0 × 107 for methanol to 3.4 × 108 M?1 s?1 for 1-octanol. Rate constants for the reactions of SO4? with deuterated methanol and ethanol are lower by about a factor of 2.5. For methanol, ethanol, and 2-propanol, the temperature dependence of the rate constant was determined over the range 10–45°C.  相似文献   

14.
The charge-transfer complex formed by the interaction of an aliphatic amine, such as n-butylamine (nBA), and carbon tetrachloride (CCl4) in dimethylsulphoxide (DMSO) initiates polymerization of methyl methacrylate (MMA) at 30°. The rate of polymerization is given by Rp = k[MMA]0.83 [nBA]0.5 [CCl4]0.5 when [CCl4]/[nBA] is ? 1. When [CCl4]/[nBA] > 1, Rp is independent of [CCl4] and Rp = k[MMA]1.46 [nBA]0.5. The average rate constants are (1.42 ± 0.05) × 10?6 1 mol?1 sec?1 in terms of MMA and (2.20 ± 0.06) × 10?6 sec?1 at 30° for higher and lower concentration of carbon tetrachloride respectively. A charge-transfer mechanism for polymerization is suggested.  相似文献   

15.
An electron paramagnetic resonance (EPR ) technique was used to show that simple alkyl radicals readily abstract hydrogen from 1,4-cyclohexadiene. Rate constants for the reaction were ca. 104–105 M?1 s?1 at 300 K and activation energies 5–7 kcal mol?1. For the stabilized radicals, allyl and benzyl, the rate constants were <102 M?1 s?1 at 300 K. The data suggest that 1,4-cyclohexadiene could be used as an effective trap to probe rearrangement reactions of carbon centered radicals and biradicals.  相似文献   

16.
It has been shown by ESR spectroscopy that the title reaction involves abstraction of hydrogen from the phosphite, since at ?10°C the reaction has a kinetic deuterium isotope effect, kH/kD, or ~3. The rate constant for hydrogen abstraction is c. 2 × 104 M?1 s?1. There is no significant addition of alkoxyl radicals to the phosphite.  相似文献   

17.
Methyl methacrylate (MMA) can be polymerized by the charge-transfer complex formed by the interaction of melamine (MM), MMA and carbon tetrachloride in a non-aqueous solvent like dimethyl sulphoxide (DMSO) or N-N-dimethylformamide. The polymerization can be accelerated by Lewis acids like Fe3?. This paper reports the polymerization of MMA initiated by MM and CCl4 and accelerated with hexakis dimethylsulphoxide iron(III) perchlorate [Fe(DMSO)6] (ClO4)3. A, at 60°. Induction periods were observed for the polymerization initiated by MM and CCl4 alone, but not when the molar ratio of MM to A became 3:1. The molecular weights of the polymers with 3:1 molar ratio of MM to A were higher than with MM alone. The rate constant for the polymerization of MMA in presence of [Fe(MM)3]3+ was 1.4181 × 10?5 1 mol?1 sec?1 at 60°. The transfer constant for CCl4, in the absence of A, is 4.66 × 10?3.  相似文献   

18.
Rate constants have been determined for the reactions of SO4? with a series of alkanes and ethers. The SO4? radical was produced by the laser-flash photolysis of persulfate, S2O82?. For methane, only an upper limit of 1 × 106 M?1 s?1 could be determined. For ethane, propane, and 2-methylpropane, rate constants of 0.44, 4.0, and 10.5 × 107 M?1 s?1 were found. For ethyl and n-propyl ether, rate constants of 1.3 × 108 and 2.2 × 108 M?1 s?1 were found and for 1,4-dioxane and tetrahydrofuran, rate constants of 7.2 × 107 and 2.8 × 108 were obtained. The reaction of SO4? with allyl alcohol was also studied and found to have a rate constant of 1.4 × 109 M?1 s?1.  相似文献   

19.
Rate constants for the gas‐phase reactions of CH3OCH2CF3 (k1), CH3OCH3 (k2), CH3OCH2CH3 (k3), and CH3CH2OCH2CH3 (k4) with NO3 radicals were determined by means of a relative rate method at 298 K. NO3 radicals were prepared by thermal decomposition of N2O5 in a 700–750 Torr N2O5/NO2/NO3/air gas mixture in a 1‐m3 temperature‐controlled chamber. The measured rate constants at 298 K were k1 = (5.3 ± 0.9) × 10?18, k2 = (1.07 ± 0.10) × 10?16, k3 = (7.81 ± 0.36) × 10?16, and k4 = (2.80 ± 0.10) × 10?15 cm3 molecule?1 s?1. Potential energy surfaces for the NO3 radical reactions were computationally explored, and the rate constants of k1k5 were calculated according to the transition state theory. The calculated values of rate constants k1k4 were in reasonable agreement with the experimentally determined values. The calculated value of k5 was compared with the estimate (k5 < 5.3 × 10?21 cm3 molecule?1 s?1) derived from the correlation between the rate constants for reactions with NO3 radicals (k1k4) and the corresponding rate constants for reactions with OH radicals. We estimated the tropospheric lifetimes of CH3OCH2CF3 and CHF2CF2OCH2CF3 to be 240 and >2.4 × 105 years, respectively, with respect to reaction with NO3 radicals. The tropospheric lifetimes of these compounds are much shorter with respect to the OH reaction. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 490–497, 2009  相似文献   

20.
Rate constants for the reactions of OH and NO3 radicals with CH2?CHF (k1 and k4), CH2?CF2 (k2 and k5), and CHF?CF2 (k3 and k6) were determined by means of a relative rate method. The rate constants for OH radical reactions at 253–328 K were k1 = (1.20 ± 0.37) × 10?12 exp[(410 ± 90)/T], k2 = (1.51 ± 0.37) × 10?12 exp[(190 ± 70)/T], and k3 = (2.53 ± 0.60) × 10?12 exp[(340 ± 70)/T] cm3 molecule?1 s?1. The rate constants for NO3 radical reactions at 298 K were k4 = (1.78 ± 0.12) × 10?16 (CH2?CHF), k5 = (1.23 ± 0.02) × 10?16 (CH2?CF2), and k6 = (1.86 ± 0.09) × 10?16 (CHF?CF2) cm3 molecule?1 s?1. The rate constants for O3 reactions with CH2?CHF (k7), CH2?CF2 (k8), and CHF?CF2 (k9) were determined by means of an absolute rate method: k7 = (1.52 ± 0.22) × 10?15 exp[?(2280 ± 40)/T], k8 = (4.91 ± 2.30) × 10?16 exp[?(3360 ± 130)/T], and k9 = (5.70 ± 4.04) × 10?16 exp[?(2580 ± 200)/T] cm3 molecule?1 s?1 at 236–308 K. The errors reported are ±2 standard deviations and represent precision only. The tropospheric lifetimes of CH2?CHF, CH2?CF2, and CHF?CF2 with respect to reaction with OH radicals, NO3 radicals, and O3 were calculated to be 2.3, 4.4, and 1.6 days, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 619–628, 2010  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