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.
In the example of ion-radical pairs (IRPs) formed in electron-transfer quenching of triplet states of the dyes methylene blue and eosin (electron acceptors) by phthalohydrazides (electron donors) in aqueous caustic solutions, it has been found that phthalohydrazide radicals are deprotonated directly in the IRP during their lifetime by the external base (OH), and a study has been made of this reaction. The yields of radicals from the deprotonated IRPs are 2–3.8 times those from the original IRPs; this difference is related to an increase in AG of IRP recombination as a result of deprotonation. In the region of high concentrations of OH, experimentally determined dependences of the yield of radicals on [OH] deviate from the relationships calculated with an exponential model of IRP destruction; this deviation is explained by a nonstationary nature of the diffusion of OH to the IRP immediately after the IRP is formed.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 6, pp. 1291–1295, June, 1990.  相似文献   

2.
The reactions of alkane and alkene oxidation with peroxynitrous acid (HOONO) in aqueous solution-gas phase systems were studied using a modified kinetic distribution method. The rate constants of oxidation of hydrocarbons (RH) were found to be unusual bell-shaped functions of the volume ratio between liquid and gas phases in a reactor. This result, as well as the previously found proportionality of the rate constants of the gas-phase RH + HOONO and RH + OH· reactions for alkanes, alkenes, and alkylbenzenes, was quantitatively interpreted assuming the rapid equilibrium distribution of HOONO and RH between a gas and a solution, the formation of OH· radicals in the two phases, and the interaction of these radicals with RH. The rate constant of peroxynitrous acid decomposition in the gas phase and the distribution coefficient of this acid between the gas phase and solution α = (0.4–2) × 10−6 were estimated. The capacity of HOONO for partition between different phases and for generation of OH· radicals in either of these phases can be of paramount importance for understanding the mechanism of lipid membrane oxidation initiated by peroxynitrous acid.__________Translated from Kinetika i Kataliz, Vol. 46, No. 3, 2005, pp. 370–379.Original Russian Text Copyright © 2005 by Lobachev, Rudakov.  相似文献   

3.
The structure of the peroxyacetic acid (PAA) molecule and its conformational mobility under rotation about the peroxide bond was studied by ab initio and density functional methods. The free rotation is hindered by the trans-barrier of height 22.3 kJ mol–1. The equilibrium molecular structure of AcOOH (C s symmetry) is a result of intramolecular hydrogen bond. The high energy of hydrogen bonding (46 kJ mol–1 according to natural bonding orbital analysis) hampers formation of intermolecular associates of AcOOH in the gas and liquid phases. The standard enthalpies of formation for AcOOH (–353.2 kJ mol–1) and products of radical decomposition of the peroxide — AcO· (–190.2 kJ mol–1) and AcOO· (–153.4 kJ mol–1) — were determined by the G2 and G2(MP2) composite methods. The O—H and O—O bonds in the PAA molecule (bond energies are 417.8 and 202.3 kJ mol–1, respectively) are much stronger than in alkyl hydroperoxide molecules. This provides an explanation for substantial contribution of non-radical channels of the decomposition of peroxyacetic acid. The electron density distribution and gas-phase acidity of PAA were determined. The transition states of the ethylene and cyclohexene epoxidation reactions were located (E a = 71.7 and 50.9 kJ mol–1 respectively).  相似文献   

4.
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   

5.
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   

6.
The standard (po = 0.1 MPa) enthalpies of formation of 2,6-di-tert-butyl-4-methylphenol and 3,5-di-tert-butylphenol in the gaseous phase, –315.5 ± 4.4 kJ mol–1 and –312.7 ± 4.6 kJ mol–1, respectively, were derived from the standard enthalpies of combustion, in oxygen, at 298.15 K, measured by static bomb combustion calorimetry, and from the standard enthalpies of sublimation, at 298.15 K, measured by Calvet microcalorimetry. The O—H bond dissociation enthalpies in those compounds were determined in benzene by photoacoustic calorimetry, leading to the standard enthalpies of formation of the gaseous phenoxy radicals: –189 ± 8 kJ mol–1 and –154 ± 6 kJ mol–1, respectively. These results were used to calculate enthalpies of substituent redistribution reactions, which are proposed as a method to estimate new data for substituted phenols.  相似文献   

7.
Experimental data on monomolecular hydrogen transfer in the reactions of the type RC·H(CH2)nCH2R1 RCH2(CH2)nC·HR1 (n = 2—4, R and R1 are alkyl substituents) were analyzed using the parabolic model (PM). The parameters characterizing this class of reactions were calculated. Isomerization of alkyl radicals via cyclic transition states (TS) is characterized by the following energy barriers to thermoneutral reaction E e0: 53.5, 65.4, and 63.2 kJ mol–1 for the six-, five-, and seven-membered TS, respectively. The E e0 energy and the strain energy change in parallel in the series of cycloparaffins CnH2n. Density functional calculations of intramolecular hydrogen transfer in the n-butyl and n-pentyl radicals and of the bimolecular hydrogen abstraction from the ethane molecule by the ethyl radical were performed. The activation energies of the intra- and intermolecular hydrogen transfer were compared. The parameters of the PM were compared with the interatomic distances in the reaction center of the TS calculated by the density functional method.  相似文献   

8.
The parabolic model of transition state has been used to analyze the problem of why aromatic amines and nitroxyl radicals cause the cyclic mechanism of chain termination in oxidizable alcohols and olefins (where HO2· and >C(OH)02· radicals participate in chain propagation) and not in oxidizable hydrocarbons. The differences are caused by the existence of a weak triplet repulsion in transition states with the N...H...0 and 0...H...0 reaction centers, while the triplet repulsion is strong in transition states with reaction centers of the C...H...0 and C...H...N type in oxidizable hydrocarbons.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1972–1976, August, 1996.  相似文献   

9.
Summary G2 theory is shown to be reliable for calculating isodesmic and homodesmotic stabilization energies (ISE and HSE, respectively) of benzene. G2 calculations give HSE and ISE values of 92.5 and 269.1 kJ mol–1 (298 K), respectively. These agree well with the experimental HSE and ISE values of 90.5±7.2 and 268.7±6.3 kJ mol–1, respectively. We conclude that basis set superposition error corrections to the enthalpies of the homodesmotic or isodesmic reactions are not necessary in calculations of the stabilization energies of benzene using G2 theory. The calculated values of the enthalpies of formation of such molecules containing multiple bonds such as benzene ands-trans 1,3-butadiene, which are found from the enthalpies of isodesmic and homodesmotic reactions rather than of atomization reactions, demonstrate good performance of G2 theory. Estimates of theH f o value for benzene from the G2 calculated enthalpies of homodesmotic reaction (2) and isodesmic reaction (3) are 80.9 and 82.5 kJ mol–1 (298 K), respectively. These are very close to the experimentalH f o value of 82.9±0.3 kJ mol–1. TheH f o value ofs-trans 1,3-butadiene calculated using the G2 enthalpy of isodesmic reaction (4) is 110.5 kJ mol–1 and is in excellent agreement with the experimentalH f o value of 110.0±1.1 kJ mol–1.  相似文献   

10.
Orthoperiodic and orthotelluric acids, their salts MIO6H4 (M = Li, Rb, Cs) and CsH5TeO6, and dimers of the salt · acid type are calculated within density functional theory B3LYP and basis set LanL2DZ complemented by the polarizationd,p-functions. According to calculations, the salt · acid dimerization is energetically favorable for compounds MIO6H4 · H5IO6 (M = Rb, Cs) and CsIO6H4 · H6TeO6. The dimerization energy is equal to 138–146 kJ mol–1. With relatively small activation energies equal to 4 kJ mol–1 (M = Li) and 11 kJ mol–1 (M = Rb, Cs), possible is rotation of octahedron IO6 relative to the M atom in monomers of salt molecules. The proton transfer along an octahedron occurs with activation energies of 63–84 kJ mol–1. The activation energy for the proton transfer between neighboring octahedrons of the type salt · acid acid · salt equals 8–17 kJ mol–1. Quantum-chemical calculations nicely conform to x-ray diffraction and electrochemical data.  相似文献   

11.
The minimal energy paths for the nucleophilic addition of a hydride ion (H) and a fluoride ion (F) to a molecule of acetylene (A) have been calculated with the use of 3–21++G and 3–21+G double basis sets in the framework of the Hartree-Fock-Roothaan method. The values of the total energies of the reactants, transition states, and products have been refined by means of calculations with more complete basis sets [6–31++G// 3–21++G and 6–31++G*//3–21++G for reaction (1); 6–31+G*//3–21+G and 6–31++G**//3–21+G for reaction (2)] and by taking into account the correlation energy for reaction (1) in the framework of the SCEP/6–31++*//3–21++G method. It has been established that the activation energy of reaction (2) is 15.94 kJ/mole lower than that for reaction (1), that reaction (1) is exothermic, and that the enthalpy change accompanying reaction (2) is close to zero. The character of the distribution of the electron density along the minimal energy paths of both reactions has been analyzed, and the differences appearing as a result of the replacement of the soft nucleophile H by the harder nucleophile F have been ascertained. The results of the calculations have been compared with the results available in the literature for reaction (1).Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 25, No. 2, pp. 149–155, March–April, 1989.  相似文献   

12.
Combustion enthalpies of three polycyclic hydrocarbons were measured by the precision bomb calorimetry method and their enthalpies of formation in the liquid state were calculated: for pentacyclo[6.3.1.13.6.02.7.09.11]tridecene, –7713.9±3.8 and 25.8±3.8 kJ mol–1; for 10-met hylpentacyclo[6.3.1.13.6.O2,.7.09.11 It ridecene, -8348.8±3.9 and -18.7+-3.9 k] mol–1; and for 11-methylpentacyclo[6.4.1.13.6.02.7.09.12]tetradecene-10-spirocyclopropane,clopropane, –10157.9±3.4 and 38.1±3.8 kJ mol–1. The thermochemical I parameters obtained agree with calculated values as well as with experimental and calculated enthalpies of formation of some hydrocarbons that contain the same fragments as the compounds studied.The authors thank Academician O. M. Nefedov and Yu. V. Tomilov for submitting the samples and for discussion of the results obtained.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 11, pp. 2676–2678, November, 1996.  相似文献   

13.
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   

14.
Summary The kinetics of the oxidation of hypophosphite ion by platinum(IV) have been studied spectrophotometrically in alkaline medium at different temperatures. The rate increases as the pH increases and the empirical rate law applicable to the reaction is given by:-d[PtIV]/dt = k3[PtIV][H2PO2–][OH]The rate constant is 2.17×10–3 (l2 mo–2s–1) at 40.5°. The energy and entropy of activation for the reaction are 104.2 kJ mol–1 and 28.5 JK–1mol–1 respectively.  相似文献   

15.
Conclusion The rate constants for the reactions of OH., NO3 ., and SO4– . with Pd(II) ions in aqueous solutions of the acids have been measured by the pulse radiolysis method. An inner-sphere mechanism of electron transfer.occurs in such reactions.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 5, pp. 1001–1004, May, 1988.  相似文献   

16.
By using the basis 3–21 + G, the minimum-energy routes for the nucleophilic addition to the F ion to methylacetylene have been calculated within the framework of the Hartree-Fock-Roothaan method according to and against the Markovnikov rule with the formation of the 1-fluoropropenyl and 2-fluoropropenyl anions. The results have been compared with data from previous calculations of the nucleophilic addition of H and F to acetylene with the formation of vinyl and fluorovinyl anions, as well as of the nucleophilic addition of H to methylacetylene according to and against the Markovnikov rule with the formation of the 1-propenyl and 2-propenyl anions. It has been established that the reaction with H is exothermic, while the reaction with F is endothermic. The activation energies of the reactions with F are lower than the activation energies of the corresponding reactions with H. It is shown that the reactions with H have a relatively early transition state, while the reactions with F are characterized by a later transition state.Translated from Teoreticheskaya i Éxperimental'naya Khimiya, Vol. 28, No. 1, pp. 5–11, January–February, 1992.  相似文献   

17.
The C5–C10 cyclolakyl radicals have a weak light absorption in the 240–300 nm wavelength range that is due to Rydberg transition to the 3s orbital. The extinction coefficients at 250 nm are in the range of 350–900 mol–1 dm3 cm–1. At this wavelength for the C6–C10 radicals a local maximum appears. The radical decay obeys second order kinetics. The kinetic characteristics of the cyclic and linear radicals are generally similar, indicating that the rings are flexible and can easily overcome steric constraints in the termination process. Both the light absorption and decay characteristics of the cyclopentyl radical are somewhat different from those of the other radicals that are attributed to the special co-planar arrangement.  相似文献   

18.
The formation of polycyclic aromatic hydrocarbons in combustion environments is linked to resonance stabilized free radicals. Here, we investigated the reaction dynamics of ground state carbon atoms, C((3)P(j)), with vinylacetylene at two collision energies of 18.8 kJ mol(-1) and 26.4 kJ mol(-1) employing the crossed molecular beam technique leading to two resonantly stabilized free radicals. The reaction was found to be governed by indirect scattering dynamics and to proceed without an entrance barrier through a long-lived collision complex to reach the products, n- and i-C(5)H(3) isomers via tight exit transition states. The reaction pathway taken is dependent on whether the carbon atom attacks the π electron density of the double or triple bond, both routes have been compared to the reactions of atomic carbon with ethylene and acetylene. Electronic structure/statistical theory calculations determined the product branching ratio to be 2:3 between the n- and i-C(5)H(3) isomers.  相似文献   

19.
Enthalpy, activation energy, and rate constant of 9 alkyl, 3 acyl, 3 alkoxyl, and 9 peroxyl radicals with alkanethiols, benzenethiol, and L ‐cysteine are calculated. The intersection parabolas model is used for activation energy calculations. Depending on the structure of attacking radical, the activation energy of reactions with alkylthiols varies from 3 to 43 kJ mol?1 for alkyl radicals, from 7 to 9 kJ mol?1 for alkoxyl, and from 18 to 35 kJ mol?1 for peroxyl radicals. The influence of adjacent π‐bonds on activation energy is estimated. The polar effect is found in reactions of hydroxyalkyl and acyl radicals with alkylthiols. The steric effect is observed in reactions of alkyl radicals with tert‐alkylthiols. All these factors are characterized via increments of activation energy. Quantum chemical calculations of activation energy and geometry of transition state were performed for model reactions: C?H3 + CH3SH, CH3O? + CH3SH, and HO2? + CH3SH with using density functional theory and Gaussian‐98. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 284–293, 2009  相似文献   

20.
Pulse radiolysis technique has been used to investigate the reaction of oxide radical ion (O.–) with 4,6-dihydroxy-2-methyl pyrimidine (DHMP), 2,4-dimethyl-6-hydroxy pyrimidine (DMHP), 5,6-dimethyl uracil (DMU) and 6-methyl uracil (MU) in strongly alkaline medium. The second-order rate constants for the reaction of O.– with these compounds are in the range 2-5 × 108 dm3 mol–1 s–1. The transient absorption spectra obtained with DHMP have two maxima at 290 and 370 nm and with DMHP have maxima at 310 and 470 nm. The transient spectrum from DMU is characterized by its absorption maxima at 310 and 520 nm and that of MU by its single maximum at 425 nm. The intermediate species were found to react with N,N,N,N-tetramethyl-p-phenylenediamine (TMPD) with high G(TMPD.+) values ranged between 3.9 × 10–7 molJ–1 and 4.8 × 10–7 molJ–1. These radicals undergo decay by second-order kinetics (2k/ = 1.0-1.7 × 106 s–1). The reaction of O.– with the selected pyrimidines is proposed to proceed through a hydrogen abstraction from the methyl group forming allyl type radicals. These are mainly oxidizing radicals and hence readily undergo electron transfer reactions with TMPD.  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