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 199 毫秒
1.
The temperature dependences of the equilibrium constants of two chain reversible reactions in quinonediimine (quinonemonoimine)—2,5-dichlorohydroquinone systems in chlorobenzene were studied. The enthalpy of equilibrium of the reversible reaction of quinonediimine with 4-hydroxydiphenylamine was estimated from these data (ΔH = − 14.4±1.6 kJ mol−1) and a more accurate value of the N-H bond dissociation energy in the 4-anilinodiphenylaminyl radical was determined (D NH = 278.6±3.0 kJ mol−1). A chain mechanism was proposed for the reaction between quinonediimine and 2,5-dichlorohydroquinone, and the chain length was estimated (ν = 300 units) at room temperature. Processing of published data on the rate constant of the reaction of styrylperoxy radicals with 2,5-dichlorohydroquinone in the framework of the intersecting parabolas method gave the O-H bond dissociation energy in 2,5-dichlorohydroquinone: D OH = 362.4±0.9 kJ mol−1. Taking into account these data, the O-H bond dissociation energy in the 2,5-dichlorosemiquinone radical was found: D OH = 253.6±1.9 kJ mol−1. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 1661–1666, October, 2006.  相似文献   

2.
The O-H bond dissociation energies (D OH) in the molecules of 2,5-dimethylhydroquinone (1) and 2,5-di-tert-butylhydroquinone (2) and in the corresponding semiquinone radicals (5 and 8, respectively) were estimated by the method of intersecting parabolas (IP) from experimental data on the rate constants for the reactions of these compounds with N-phenyl-1,4-benzoquinonemonoimine (3) and using the density functional B3LYP/6-31+G* quantum chemical calculations. When calculating the D OH values by the IP method, solvation of reactants and transition states should be taken into account. The energies of solvation of quinones, semiquinone radicals, and hydroquinones were evaluated by the PCM method. The results of quantum chemical calculations obtained with inclusion of the effects of solvation and the D OH estimates obtained by the IP method are in good agreement, being equal to 337.9±1.6, 242.5±1.4, and 242.7±3.4 kJ mol−1 for molecule 1 and radicals 5 and 8, respectively. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2244–2251, October, 2005.  相似文献   

3.
Intramolecular 1-n H-shift (n = 2, 3… 7) reactions in alkoxy, alkyl and peroxy radicals were studied by density functional theory (DFT) at the B3LYP/6-311+G∗∗ level and compared with respective intermolecular H-transfers. It was found that starting from 1 to 3 H-shift the barrier heights stepwise decrease with increasing n reaching minimum for 1-5 and 1-6 H-shifts. This dependence can be ascribed to the decrease of the strain with increasing transition state (TS) ring size, which is minimal in six- and seven-member rings. The barrier heights of H-shifts in alkyl radicals are systematically larger than those in alkoxy radicals: the respective activation energies (Ea) of 1-5 and 1-6 H-shifts are about 59-67 kJ/mol for alkyl radical and 21-34 kJ/mol for alkoxy radicals. Further increase of the TS ring size in 1-7 H-shifts leads to the increase of the barrier to 44 kJ/mol in the hexyloxy radical and 84 kJ/mol for n-heptyl radical. We have also found that intermolecular H-transfer reactions in all three types of free radicals have smaller barriers than respective intramolecular 1-5 or 1-6 H-shifts by 4-25 kJ/mol. The mentioned difference can be explained in terms of enhanced nonbonding repulsion interaction in the cyclic TS structures compared to respective intermolecular TS. B3LYP/6-311+G∗∗ geometric parameters and imaginary frequencies for 1-n H-shifts TS are consistent with respective calculated barrier heights. Reactivity of some other radicals compared to alkoxy, peroxy and alkyl radicals as well as other factors influencing their reactivity (π-conjugation, steric effect and ring strain in cyclic TS, etc.) are also briefly discussed in relation to free radical reactions in polymer systems.  相似文献   

4.
The dissociation energies of O-H bonds in natural antioxidants were estimated from the kinetic data (rate constants of reactions of the antioxidants with the peroxy radicals). The calculations were performed using the method of intersecting parabolas. α-Tocopherol was used as a reference phenol with D O-H = 330 kJ mol−1. The following groups of antioxidants were chosen for the estimation: tocopherols, their sulfur- and selenium-containing analogs, flavones, flavanones, and gallates. The discrepancy in the D O-H values for the same phenol by measurements of k(RO2 · + ArOH) of different authors does not exceed 2 kJ mol−1. The D O-H values were calculated for 64 phenols. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1824–1832, September, 2008.  相似文献   

5.
The O-H bond dissociation energies (D O-H) in five alcohols and six acids have been determined from experimental data (rate constants of radical reactions). The ratio of the rate constants of the reactions R1O˙+RH→R1OH+R˙ and R i O˙+RH→R i OH+R˙ and the intersecting parabolas method are used in the estimation procedure. The D O-H values are used to calculate the activation energies and rate constants for hydrogen abstraction from 2-methylbutane, butene-1, and cumene by alkoxyl and carboxyl radicals. The geometric parameters of the transition state are calculated for these reactions.  相似文献   

6.
The rate constants and activation parameters for the self-coupling of perfluoroperoxy radicals of structure A and B: C7F15OO (A) and RFOCF2OO (B) have been determined in perfluorohexane solution in the temperature range 228-258 K. The magnitude of the rate constants obtained ranks between 6.6×108 and 2.5×109 l mol−1 s−1 and are therefore, among the largest rate values so far reported in the literature for primary peroxy radicals couplings. The activation energy is positive and lower for the peroxy radicals (A) with respect to the peroxy radicals (B) (10.5 and 23.0 kJ mol−1, respectively).Analysis by kinetic modeling has shown that the peroxy radicals decay curves are compatible with the participation of peroxy radicals↔tetroxide equilibria to the reaction mechanism.Upper limit values of kbs<10 and <20 s−1 were inferred for the β-scission reactions of the perfluoroalkylperoxy radicals at 228 and 258 K, respectively.  相似文献   

7.
The transesterification of methyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl)propanoate with tetrakis(hydroxymethyl)methane depends on the equilibrium constants of the reversible reactions; for the final step, the equilibrium constant is K ? 1. The molecular geometries and the enthalpies and entropies of the equilibrium reactions were calculated by the semiempirical PM6 quantum chemical method. The thermodynamic equilibrium constants of the reversible reactions were calculated by the Boltzmann equation from the Gibbs energies G f . For tris-[3-(3,5-di-tert-butyl-4-hydroxyphenyl)propanoyloxymethyl](hydroxymethyl)methane, the dipole moment is μ = 0.97 D and the energy of the O-H homolysis is D OH = 347.3 kJ mol?1. For tetrakis[3-(3,5-di-tert-butyl-4-hydroxyphenyl)propanoyloxymethyl]methane, μ is 5.6 D and D OH is 321 kJ mol?1. The geometry of the structure affects the H-O homolysis energy and the chain termination coefficient under the conditions of inhibited cumene oxidation.  相似文献   

8.
The enthalpies of intramolecular reactions of alkoxy and peroxy radicals formed from polyatomic artemisinin hydroperoxides and of their bimolecular reactions with C—H, S—H, and O—H bonds of biological substrates were calculated. The activation energies and rate constants of these reactions were calculated using the intersecting parabolas method. The decomposition of artemisinin hydroperoxides can initiate the cascade of intramolecular oxidation reactions involving radicals R·, RO·, HO·, HO2·, and RO2·. The main sequences of transformation of these radicals were established. The oxidative destruction of the artemisinin peroxy derivatives generates radicals RO2·, HO·, and HO2· in an amount of 4.5 radicals per peroxide derivative molecule on the average. The kinetic scheme of oxidative transformations of the hydroperoxide with four OOH groups and radicals formed from it was constructed using this radical as an example.  相似文献   

9.
Hydroperoxyl (HOO.) and alkylperoxyl (ROO.) radicals show a different behavior in H‐atom‐transfer processes. Both radicals react with an analogue of α‐tocopherol (TOH), but HOO., unlike ROO., is able to regenerate TOH by a fast H‐atom transfer: TO.+HOO.→TOH+O2. The kinetic solvent effect on the H‐atom transfer from TOH to HOO. is much stronger than that observed for ROO. because noncovalent interactions with polar solvents (Solv???HOO.) destabilize the transition state.  相似文献   

10.
A large body of experimental data on the reactions of ozone with C-H bonds of polar molecules in the liquid and gas phases is analyzed in the framework of the intersecting parabolas model. The reactions are considered as the abstraction reaction O3 + RH → HOOO. + R.. The contribution from the polar effect to the activation energy of such reactions is calculated. This contribution is ?6.8 kJ/mol for the reactions of ozone with aliphatic alcohols, and is ?8.1, ?11.7, ?6.8, and ?2.2 kJ/mol for the reactions of ozone with ketones, ethers, 1,3-dioxolanes, and 1,3-dioxanes, respectively. The contribution is insignificant in the reactions of ozone with aldehydes. The interatomic distances in the transition state of these reactions r #(C…H) and r #(O…H) and the angle between the C…H and O…H bonds are calculated. For the reactions in polar solvents, the contribution from solvation to the activation energy is calculated. In most of the systems considered, this contribution is insignificant (from ?1 to ?3 kJ/mol). The reactions involving ozone are compared to the reactions of peroxy radicals with the same classes of compounds.  相似文献   

11.
The parabolic model of radical abstraction reactions is used to analyze experimental data on monomolecular hydrogen-atom transfer in the reactionsRC.H(CH2) n CH2R1 RCH2(CH2) n C.HR1(n= 2, 3, 4)RCH(O.)(CH2)2CH2R1 RCH(OH)(CH2)2C.HR1 RCH(OO.)(CH2) n CH2R1 RCH(OOH)(CH2) n C.HR1(n= 1, 2).The activation energies and rate constants that specify each class of these reactions are calculated. Alkyl radical isomerization is characterized by the following activation energies of a thermally neutral reaction depending on the cycle size in the transition state (nis the number of atoms in a cycle): E e , 0(kJ/mol) = 46.6 (n= 6), 59.4 (n= 5), and 57.1 (n= 7). Alkoxy radicals isomerize with E e , 0(kJ/mol) = 53.4 (n= 6), whereas peroxy radicals isomerize with E e , 0(kJ/mol) = 53.2 (n= 6) and E e , 0(kJ/mol) = 54.8 (n= 7). The E e , 0value varies with changes in the cycle size and the strain energy in cycloparaffin C n H2n in the same manner. The activation energies E e , 0for the intra- and intermolecular H-atom abstractions are compared. It is found that E e , 0(isomerization) < E e , 0(R.+ R1H) for alkyl radicals and that E e , 0(isomerization) E e , 0(RO.(RO.) + R1H) for alkoxy and peroxy radicals.  相似文献   

12.
The rate constants of addition of the.CMe3,.CH2Me,.CH2(CH2)3Me,.CH2Ph,.CH2CH=CH2, and.CH(Me)Et radicals to fullerene C60 were determined by the method of competitive addition of free radicals to spin traps. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp 2369–2372, December, 1999.  相似文献   

13.
The dissociation energies of O-H and N-H bonds have been determined for ten aminophenoltype (HOArAmH) hybrid antioxidants. The bond dissociation energies D O-H and D N-H have been estimated from experimental kinetic data (rate constants of the reactions of peroxyl radicals with these antioxidants and their alkyl-substituted derivatives) by the intersecting-parabolas method. Kinetic data for the reactions of peroxyl radicals with HOArAmH, ROArAmH, and HOArAmR compounds were used. The following D O-H and D N-H values (kJ/mol) were obtained: for 4-hydroxydiphenylamine, D O-H = 338.8 and D N-H = 355.9; for 4-hydroxyphenyl-2-naphthylamine, D O-H = 335.4 and D N-H = 353.6; for 6-hydroxy-1,2-dihydro-2,2,4-tri-methylquinoline, D O-H = 338.0 and D N-H = 348.2; for 9-hydroxy-1,2-dihydro-2,3,4-trimethylquinoline, D O-H = 329.7 and D N-H = 383.3; for 6-hydroxy-1,2,3,4-tetrahydro-2,2,4-trimethylquinoline, D O-H = 324.4 and D N-H = 345.3; for 8-hydroxy-1,2,3,4-tetrahydro-2,2,4-trimethylquinoline, D O-H = 329.4 and D N-H = 380.6; for 5-hydroxyimidazole, D O-H = 356.4 and D N-H = 368.4; for 5-hydroxy-2-methylimidazole, D O-H = 351.3 and D N-H = 362.6; for 5-hydroxy-4,6-dimethylimidazole, D O-H = 346.7 and D N-H = 357.3; for 5-hydroxy-2,4,6-trimethylimidazole, D O-H = 347.7 and D N-H = 358.7.  相似文献   

14.
The N—H and O—H bond dissociation energies in 4-hydroxydiphenylamine Ph—NH—C6H4—OH (D NH= 353.4, D OH=339.3 kJ mol–1) and its semiquinone radicals D NH(Ph—NH—C6H4—O·) = 273.6, D OH(Ph—N·—C6H4—OH) = 259.5 kJ mol–1 were first estimated using the parabolic model and experimental data (rate constants) on two elementary reactions with participation of N-phenyl-1,4-benzoquinonemonoimine (2). One of the reactions, namely, that of 2 with aromatic amines, was studied in this work using a specially developed method.  相似文献   

15.
The reactivity of RO 2 . peroxide radicals in oxidation reactions depends greatly on specific and nonspecific solvation by the solvent. With increasing dielectric constant () of the medium, the rate constant for the interaction of RO 2 . radicals with methyl ethyl ketone (k2) and the rate constant of the recombination of RO 2 . radicals (k6) increase. The specific solvation of RO 2 . radicals due to hydrogen bonds of water diminishes their reactivity. Equilibrium constants for the solvation of peroxide radicals and all rate constants of chain propagation and termination reactions involving solvated and unsolvated RO 2 . radicals were measured. The change in composition of the oxidation products when methyl ethyl ketone was diluted with benzene or water is caused by nonspecific and specific solvation by the solvent affecting the chain propagation reaction.  相似文献   

16.
An ESR method for studying the mechanism of H-transfer reactions between H-donors of different reactivity (A1H, A2H…) and their free radicals (A1; A2.…) in non-polar solvents at ambient temperature is presented. The new technique is based on a pulsed initiation of various secondary phenoxy or nitroxy radicals in binary mixtures of hindered phenols, unhindered phenols, partially hindered thiobisphenols and diphenylamine, employing a high concentration of free RO2. and coordinated (CoIII)RO2. tert-butyl peroxy radicals generated in the redox-reaction of Co(acac)2 with tert-butyl hydroperoxide. The consecutive H-transfer reactions proceed to equilibrium until the most stable radicals are formed. In this way criteria are obtained for ranking the compared free and coordinated phenoxy radicals according to their relative stabilities. The secondarily generated phenoxy radicals from unhindered phenols after coordination to CoIII are stabilized and cannot take part in further H-transfer reactions.  相似文献   

17.
The data on enthalpies of formation (Δf H ) of alkylcarbonyl radicals are expanded to 22 items. The reference value of Δf H for diacetylperoxide in the gas phase (?485.5 kJ/mol) is determined via recalculation from evaporation enthalpy (n-C5H11C(O)O)2. The Δf H values of 22 diacylperoxides (gaseous) are calculated and used in combination with the literature data on dissociation energies of D(O-O) bonds in them to determine the Δf H of corresponding radicals. The interrelation between structure and properties (the enthalpy of formation) is considered, and the parameters for the calculated prediction of Δf H are found.  相似文献   

18.
Conclusions The rate constants of the reaction between the alkyl radicals of hydrocarbons (R.) and of vinyl monomers (M.) and O2 and stable nitroxyl radicals were determined at 50C. The low-molecular-weight radicals react with O2 (k1) and the (k3) more rapidly than the M. do. For polar R. and M. k1 and k3 are close, and for the nonpolar ones k1>k3.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 11, pp. 2446–2450, November, 1979.The authors express their appreciation to V. A. Golubev for submitting the nitroxyl radicals and A. P. Moravskii for help in running the experiment.  相似文献   

19.
A method to determine the relative antioxidant capacity of radical scavengers based on the inhibition of the oscillations of the Briggs? Rauscher (BR) oscillating reaction was previously reported. A semiquantitative mechanistic interpretation of the inhibitory effects required two steps to obtain simulated inhibition times in very good agreement with the experimental ones. The first step is inhibitory, involving H‐atom transfer from antioxidant to the HOO. radical; the second step is a first‐order degradation of the antioxidant to unspecified products. Since the degradation may be due to oxidation and/or iodination of the antioxidant, we studied the kinetics of the subsystems IO (H+)+antioxidant and I2(H+)+antioxidant. We used 2,5‐ and 2,6‐dihydroxybenzoic acids, caffeic acid (=3‐(3,4‐dihydroxyphenyl)prop‐2‐enoic acid), ferulic acid (=3‐(4‐hydroxy‐3‐methoxyphenyl)prop‐2‐enoic acid), pyrocatechol (=benzene‐1,2‐diol), and hydroquinone (=benzene‐1,4‐diol) as antioxidants. Spectra in the wavelength range 500–250 nm were repeated at given time intervals to follow the peaks of the iodine and oxidation products, which were mainly quinones. For the iodination of the above diphenols (=benzenediol derivatives) the substitution and/or addition reactions with I2 or HOI were found to be relatively slow compared to oxidation by IO . Approximate rate constants for oxidation were obtained on the basis of a reasonable kinetic model by using a suitable numerical integration program. Although these complexities can arise also in the completely inhibited BR oscillator, we believe that the inhibitory effects are due to the HOO. scavenging action by diphenols or by quinones since HOO. radicals are also potential reducing agents. We propose two steps that could maintain a small reservoir of diphenol, while both quinone and diphenol deplete HOO. radicals. In short, the complexities do not affect the method for monitoring the relative activity of antioxidants based on the BR oscillating reaction. The effects of temperature on the inverse of the oscillatory time in the BR‐uninhibited system, on the inverse of inhibition times, and on the time length of the resumed oscillations for four antioxidants were also investigated. Apparent average activation energies were obtained.  相似文献   

20.
Gas-phase reactions typical of the Earth’s atmosphere have been studied for a number of partially fluorinated alcohols (PFAs). The rate constants of the reactions of CF3CH2OH, CH2FCH2OH, and CHF2CH2OH with fluorine atoms have been determined by the relative measurement method. The rate constant for CF3CH2OH has been measured in the temperature range 258–358 K (k = (3.4 ± 2.0) × 1013exp(?E/RT) cm3 mol?1 s?1, where E = ?(1.5 ± 1.3) kJ/mol). The rate constants for CH2FCH2OH and CHF2CH2OH have been determined at room temperature to be (8.3 ± 2.9) × 1013 (T = 295 K) and (6.4 ± 0.6) × 1013 (T = 296 K) cm3 mol?1 s?1, respectively. The rate constants of the reactions between dioxygen and primary radicals resulting from PFA + F reactions have been determined by the relative measurement method. The reaction between O2 and the radicals of the general formula C2H2F3O (CF3CH2? and CF3?HOH) have been investigated in the temperature range 258–358 K to obtain k = (3.8 ± 2.0) × 108exp(?E/RT) cm3 mol?1 s?1, where E = ?(10.2 ± 1.5) kJ/mol. For the reaction between O2 and the radicals of the general formula C2H4FO (? HFCH2O, CH2F?HOH, and CH2FCH2?) at T = 258–358 K, k = (1.3 ± 0.6) × 1011exp(?E/RT) cm3 mol?1 s?1, where E = ?(5.3 ± 1.4) kJ/mol. The rate constant of the reaction between O2 and the radicals with the general formula C2H3F2O (?F2CH2O, CHF2?HOH, and CHF2CH2?) at T = 300 K is k = 1.32 × 1011 cm3 mol?1 s?1. For the reaction between NO and the primary radicals with the general formula C2H2F3O (CF3CH2? and CF3?HOH), which result from the reaction CF3CH2OH + F, the rate constant at 298 K is k = 9.7 × 109 cm3 mol?1 s?1. The experiments were carried out in a flow reactor, and the reaction mixture was analyzed mass-spectrometrically. A mechanism based on the results of our studies and on the literature data has been suggested for the atmospheric degradation of PFAs.  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