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.
M?ller-Plesset perturbation theory and density functional theory calculations are used to study decomposition mechanisms of polymerization initiators, such as diethyl peroxydicarbonate, trifluoroacetyl peroxide, and acetyl peroxide, which possess a general structure of RC(O)OO(O)CR. It is found that the decomposition of initiators with electron-donating R groups follows two favorable stepwise pathways: a two-bond cleavage mechanism in which the O-O single bond and one of R-C bonds of [R-C(O)O-O(O)C-R] break simultaneously followed by decomposition of the R-C(O)O(*) radical and a one-bond cleavage mechanism in which the single O-O bond cleavage produces a carboxyl radical pair and a subsequent decomposition of the carboxyl radicals. It is also found that the initiators with electron-withdrawing R groups follow the two-bond cleavage pathway only. Geometrical and energetic analyses indicate that despite the similar structures of the peroxydicarbonates, quite different decomposition energy barriers are determined by the nature of the R groups.  相似文献   

2.
3.
A novel fullerenyl cation (EtO)2P+(OH)CH2-C60+ was generated by simply dissolving the monofunctionalized hydrofullerene RC60-H or singly bonded dimer RC60-C60R (R = CH2P(O)(OEt)2) in oxidizing acids such as H2SO4 and FSO3H. The cation was also formed in CH2Cl2 by one-electron oxidation with aminium radical cation and was used for further functionalization of C60.  相似文献   

4.
The copper-catalyzed radical cyclization (RC) of N-α-perchloroacyl cyclic ketene-N,X(X=O, NR, S)-acetals was studied. While the RC of N-acyl ketene-N,O-acetals was unsuccessful, the 5-endo cyclization of the other ketene acetals provided much better results, with the following order of cyclization efficiency: hexa-atomic cyclic ketene-N,NR-acetals<penta-atomic cyclic ketene-N,S-acetals<hexa-atomic cyclic ketene-N,S-acetals. Invariably the catalytic cycle begins with the formation of a carbamoyl methyl radical. This leads to a cascade of reactions, including a radical polar crossover step, which ends with the formation of the maleimide nucleus, or precursors of this. Products from the RC of the hexa-atomic cyclic ketene-N,S-acetals, were efficiently transformed into disubstituted maleic anhydrides.  相似文献   

5.
Abstract— By means of in situ photolysis EPR of aqueous solutions of α-oxocarboxylic acids (RCO-CO2H) at pH values above 5, semidione radical anions [RC(O-)=C(O')R] and α-hydroxy-α-carboxy alkyl radicals [RC(OH)CO2-] were detected. C02 was identified as a reaction product. On photolysis of mixtures of α-oxocarboxylic acids (RCOCO2H and R'COCC2H), "mixed" semidione radical anions [RC(O->=C(O)R'] were observed in addition to RC(O-)=C(O')R, R'C(O-)=C(O')R', RC(OH)CO2- and R'C(OH)CO2-. The experimental results are explained in terms of photodecarboxylation (α-clea-vage) of electronically excited RCOCOJ to yield RCO and CO2. The radicals RC(OH)CO2- are formed by reduction of RCOCO2- by CO2-. The semidione radicals are produced by addition of RCO to RCOCO2- followed by decarboxylation of the intermediate adduct. This mechanism was confirmed by generating acyl radicals independently and reacting them with α-oxocarboxylic acids. Selected product studies support the mechanism suggested.  相似文献   

6.
The Cl‐atom‐initiated oxidation of two esters, ethyl formate [HC(O)OCH2CH3] and ethyl acetate [CH3C(O)OCH2CH3], has been studied at pressures close to 1 atm as a function of temperature (249–325 K) and O2 partial pressure (50–700 Torr), using an environmental chamber technique. In both cases, Cl‐atom attack at the CH2 group is most important, leading in part to the formation of radicals of the type RC(O)OCH(O?)CH3 [R = H, CH3]. The atmospheric fate of these radicals involves competition between reaction with O2 to produce an anhydride compound, RC(O)OC(O)CH3, and the so‐called α‐ester rearrangement that produces an organic acid, RC(O)OH, and an acetyl radical, CH3C(O). For both species studied, the α‐ester rearrangement is found to dominate in air at 1 atm and 298 K. Barriers to the rearrangement of 7.7 ± 1.5 and 8.4 ± 1.5 kcal/mole are estimated for CH3C(O)OCH(O?)CH3 and HC(O)OCH(O?)CH3, respectively, leading to increased occurrence of the O2 reaction at reduced temperature. The data are combined with those obtained from similar studies of other simple esters to provide a correlation between the rate of occurrence of the α‐ester rearrangement and the structure of the reacting radical. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 397–413, 2010  相似文献   

7.
陈文祥  Allen  N.S 《化学学报》1990,48(6):608-611
用闪光光介研究的结果显示了水溶性甲磺酸二苯甲酮季铵盐(1)经由n-λ^*激发导致三线态, 与叔胺形成三线态激发复合物, 同时再从叔胺夺取一个电子产生自由基阴离子。经过分子间抽氢产生ketyl自由基和烷基自由基, 通过对氧、PH和胺电离势的影响的研究证实了上述机理。  相似文献   

8.
When the superoxide radical O(2)(?-) is generated on reaction of KO(2) with water in dimethyl sulfoxide, the decay of the radical is dramatically accelerated by inclusion of quinones in the reaction mix. For quinones with no or short hydrophobic tails, the radical product is a semiquinone at much lower yield, likely indicating reduction of quinone by superoxide and loss of most of the semiquinone product by disproportionation. In the presence of ubiquinone-10, a different species (I) is generated, which has the EPR spectrum of superoxide radical. However, pulsed EPR shows spin interaction with protons in fully deuterated solvent, indicating close proximity to the ubinquinone-10. We discuss the nature of species I, and possible roles in the physiological reactions through which ubisemiquinone generates superoxide by reduction of O(2) through bypass reactions in electron transfer chains.  相似文献   

9.
用亚硝基丁烷(TNB)、亚硝基苯和C-苯基-N-叔丁基硝酮作为自旋捕获剂来捕获和检出一些二酰基过氧化物类化合物分解时形成的短命烷基自由基. 报导了烷基自由基对自旋捕获剂加成的氮氧化物自由基的电子自旋共振谱, 并观察到2,4,4-三甲基戊基自由基中α-CH2的氢是非对映和非等价的. 据此, 认为与这一类自由基的加合物产生的电子自旋共振谱与其它类型自由基是不同的.  相似文献   

10.
In the present work the potential energy surface (PES) corresponding to the different initiation routes of the oxidation mechanism of DMS by hydroxyl radical in the absence of O(2) has been studied, and connections among the different stationary points have been established. Single-point high level electronic structure calculations at lower level optimized geometries have been shown to be necessary to assure convergence of energy barriers and reaction energies. Our results demonstrate that the oxidation of DMS by OH turns out to be initiated via three channels: a hydrogen abstraction channel that through a saddle point structure finally leads to CH(3)SCH(2) + H(2)O, an addition-elimination channel that firstly leads to an adduct complex (AD) and then via an elimination saddle point structure finally gives CH(3)SOH and CH(3) products, and a third channel that through a concerted pathway leads to CH(3)OH and CH(3)S. The H-abstraction and the addition-elimination channels initiate by a common pathway that goes through the same reactant complex (RC). Our theoretical results agree quite well with the branching ratios experimentally assigned to the formation of the different products. Finally, the calculated equilibrium constants of the formation of the complex AD and the hexadeuterated complex AD from the corresponding reactants, as a function of the temperature, are in good accordance with the experimental values.  相似文献   

11.
The mechanism of hydroxyl radical initiated degradation of a typical oil sands process water (OSPW) alicyclic carboxylic acid was studied using cyclohexanoic acid (CHA) as a model compound. By use of vacuum ultraviolet irradiation (VUV, 172 nm) and ultraviolet irradiation in the presence of hydrogen peroxide UV(254 nm)/H(2)O(2), it was established that CHA undergoes degradation through a peroxyl radical. In both processes the decay of the peroxyl radical leads predominantly to the formation of 4-oxo-CHA, and minor amounts of hydroxy-CHA (detected only in UV/H(2)O(2)). In UV/H(2)O(2), additional 4-oxo-CHA may also have been formed by direct reaction of the oxyl radical with H(2)O(2). The oxyl radical can be formed during decay of the peroxyl-CHA radical or reaction of hydroxy-CHA with hydroxyl radical. Oxo- and hydroxy-CHA further degraded to various dihydroxy-CHAs. Scission of the cyclohexane ring was also observed, on the basis of the observation of acyclic byproducts including heptadioic acid and various short-chain carboxylic acids. Overall, the hydroxyl radical induced degradation of CHA proceeded through several steps, involving more than one hydroxyl radical reaction, thus efficiency of the UV/H(2)O(2) reaction will depend on the rate of generation of hydroxyl radical throughout the process. In real applications to OSPW, concentrations of H(2)O(2) will need to be carefully optimized and the environmental fate and effects of the various degradation products of naphthenic acids considered.  相似文献   

12.
The EPR-spectral zero-field splitting parameter D of the localized heterovinyl-substituted 1,3-cyclopentanediyl triplet diradicals T, generated in a 2-methyltetrahydrofuran (MTHF) glass matrix at 77 K through the photochemical deazetation of the corresponding azoalkanes 1-13, is a quantitative measure of the spin-density (rho) variation by the substituents at the radical site in the 1-heteroallylic radicals. From these data, the radical-stabilizing ability of a variety of nitrogen-containing groups has been assessed, which includes imino and hydrazonyl functionalities. The radical stabilization in the heteroallylic radical fragment follows the order X = O < NMe < CH2 < CHMe < NOH approximately NOMe < NNHCHO approximately NNHC(O)NH2 < NPh approximately NNMe2 < NNH2 < CHPh < NNHPh. The lowest D values have been found for the hydrazonyl-substituted derivatives, which implies the lowest spin density at the carbon center and, thus, the most efficacious radical stabilization through spin delocalization. This superdelocalization may be rationalized in terms of nitrogen-centered radical-cationic structures. Localization of the spin at the terminal atom is resisted through the electronegativity effect (O < N < C).  相似文献   

13.
It has been established that reversible photoinduced transformations of 2,3-dimethyloxirane and methyloxirane radical cations (RCs), observed in freonic matrices at 77 K, are related to the conversion between the open and cyclic forms of the RCs. For the trimethyloxirane RC the action of light on the trans-isomer of the open form results in its photoinduced transformation into a C-centered radical with low quantum efficiency (≈4 × 10−3). Upon the X-ray irradiation of 2,2-dimethyloxirane in freonic matrices at 77 K, a cyclic form of the RC is stabilized (presumably, as part of a complex with matrix molecules) which transforms into a distonic C-centered RC under the action of light with the quantum yield of ≈10−3. Tetramethyloxirane RC, stabilized in its open form, is resistant to the action of light. Probable causes of the observed effects are discussed.  相似文献   

14.
In the reaction of TAS-fluoride, (Me2N)3S+Me3SiF2-, with carbonyl sulfur difluoride imides RC(O)NSF2 (R = F, CF3), C-N bond, cleavage is observed, and TAS+RC(O)F2- and NSF are the final products. From TASF and RC(O)NS(CF3)F, the salts TAS+RC(O)NS(CF3)F2- (R = F (14), CF3 (15)), with psi-pentacoordinate sulfur centers in the anions, are formed. An X-ray structure investigation of 14 shows that the fluorine atoms occupy axial positions and CF3, NC(O)F, and the sulfur lone pair occupy equatorial positions of the trigonal bipyramid. The -C(O)F group lies in the equatorial plane with the CO bond synperiplanar to the SN bond. According to B3LYP calculations, this structure corresponds to a global minimum and the expected axial orientation of the -C(O)F group represents a transition state. Calculations for the unstable FC(O)NSF3- anion show a different geometry. The -C(O)F group deviates 40 degrees from axial orientation, and the equatorially bonded fluorine is, in contrast to the -CF3 group in 14, syn positioned.  相似文献   

15.
MIVDO/3方法研究RC(X)OC(O)CH3热分解反应   总被引:1,自引:0,他引:1  
用MINDO/3方法研究了酸酐(X=O)和硫代酸酐(X=S)的热分解反应。结果表明,含有β-氢原子的乙酸酐和硫代乙酸酐可按三条路径分解,苯基对热分解的影响不大。  相似文献   

16.
The review briefly outlines theoretical models developed in 1990s to describe electron transfer reactions (ETR) in proteins, as well as different variants of improvements in these models proposed by the present authors to describe ETR in reaction centers (RC) of photosynthetic bacteria with consideration of their molecular dynamics in a wide temperature range. Experimental data on electron transfer from reduced proximal heme c-559 of cytochrome to bacteriochlorophyll dimer radical cation P+ in RC from two types of bacteria, viz., native and mutant RC from Rps. viridis and native RC from Rps. sulfoviridis were analyzed within the framework of the models which take into account the quantum and classical (including diffusive) degrees of freedom responsible for reorganization of the protein globule.  相似文献   

17.
Aldehydes are important intermediates and products in a variety of combustion and gas-phase oxidation processes, such as in low-temperature combustion, in the atmosphere, and in interstellar media. Despite their importance, the enthalpies of formation and bond dissociation energies (BDEs) for the aldehydes are not accurately known. We have determined enthalpies of formation for acetaldehyde, propanal, and butanal from thermodynamic cycles, using experimentally measured reaction and formation enthalpies. All enthalpy values used for reference molecules and reactions were first verified to be accurate to within around 1 kcal mol-1 using high-level ab initio calculations. Enthalpies of formation were found to be -39.72 +/- 0.16 kcal mol-1 for acetaldehyde, -45.18 +/- 1.1 kcal mol-1 for propanal, and -49.27 +/- 0.16 kcal mol-1 for butanal. Enthalpies of formation for these three aldehydes, as well as for pentanal, hexanal, and heptanal, were calculated using the G3, G3B3, and CBS-APNO theoretical methods, in conjunction with bond-isodesmic work reactions. On the basis of the results of our thermodynamic cycles, theoretical calculations using isodesmic work reactions, and existing experimental measurements, we suggest that the best available formation enthalpies for the aldehydes acetaldehyde, propanal, butanal, pentanal, hexanal, and heptanal are -39.72, -45.18, -50.0, -54.61, -59.37, and -64.2 kcal mol-1, respectively. Our calculations also identify that the literature enthalpy of formation of crotonaldehyde is in error by as much as 1 kcal mol-1, and we suggest a value of -25.1 kcal mol-1, which we calculate using isodesmic work reactions. Bond energies for each of the bonds in the aldehydes up to pentanal were calculated at the CBS-APNO level. Analysis of the BDEs reveals the R-CH(2)CH=O to be the weakest bond in all aldehydes larger than acetaldehyde, due to formation of the resonantly stabilized vinoxy radical (vinyloxy radical/formyl methyl radical). It is proposed that the vinoxy radical as well as the more commonly considered formyl and acetyl radicals are important products of aldehyde combustion and oxidation, and the reaction pathways of the vinoxy, formyl, and acetyl radicals are discussed. Group additivity values for the carbon-oxygen-hydrogen groups common to the aldehydes are also determined. Internal rotor profiles and electrostatic potential surfaces are used to study the dipole induced dipole-dipole interaction in the synperiplanar conformation of propanal. It is proposed that the loss of this dipole-dipole interaction in RC(.-)HCH(2)CH=O radicals causes a ca. 1-2 kcal mol-1 decrease in the aldehyde C-H and C-C bond energies corresponding to RC(.-)HCH(2)CH=O radical formation.  相似文献   

18.
Among various types of radical reactions, the addition of carbon radicals to unsaturated bonds is a powerful tool for constructing new chemical bonds, in which the typical applied unsaturated substrates include alkenes, alkynes and imines. Carbonyl is perhaps the most common unsaturated group in nature. This work demonstrates a novel C?O bond formation through carbon‐centered radical addition to the carbonyl oxygen of amide or ester, in which amide and ester groups are easily activated through the radical process. EPR spectroscopy and radical clock experiments support the radical process for this transformation, and density functional theory (DFT) calculations support the possibility of carbon‐centered radical addition to the carbonyl oxygen of amides or esters.  相似文献   

19.
We present the results of a reactivity study of titanium cationic clusters towards CO, C(2)H(2), C(2)H(4) and C(3)H(6) based on guided-ion-beam mass spectrometry and DFT calculations. We identified Ti(2)O(4)(+) and to a lesser extent TiO(2)(+) species which preferentially undergo oxidation reactions. An oxygen centered radical of Ti(2)O(4)(+) is responsible for selective oxidation. Energy profiles and MD simulations reveal the mechanisms of the reactions. Regeneration of the oxygen centered radical was achieved experimentally and theoretically through the reaction of N(2)O with Ti(2)O(3)(+).  相似文献   

20.
Pulse radiolysis and flash photolysis are used to generate the hyponitrite radicals (HN2O2(*)/N2O2(*-)) by one-electron oxidation of the hyponitrite in aqueous solution. Although the radical decay conforms to simple second-order kinetics, its mechanism is complex, comprising a short chain of NO release-consumption steps. In the first, rate-determining step, two N2O2(*-) radicals disproportionate with the rate constant 2k = (8.2 +/- 0.5) x 10(7) M(-1) s(-1) (at zero ionic strength) effectively in a redox reaction regenerating N2O2(2-) and releasing two NO. This occurs either by electron transfer or, more likely, through radical recombination-dissociation. Each NO so-produced rapidly adds to another N2O2(*-), yielding the N3O3(-) ion, which slowly decomposes at 300 s(-1) to the final N2O + NO2(-) products. The N2O2(*-) radical protonates with pKa = 5.6 +/- 0.3. The neutral HN2O2(*) radical decays by an analogous mechanism but much more rapidly with the apparent second-order rate constant 2k = (1.1 +/- 0.1) x 10(9) M(-1) s(-1). The N2O2(*-) radical shows surprisingly low reactivity toward O2 and O2(*-), with the corresponding rate constants below 1 x 10(6) and 5 x 10(7) M(-1) s(-1). The previously reported rapid dissociation of N2O2(*-) into N2O and O(*-) does not occur. The thermochemistry of HN2O2(*)/N2O2(*-) is discussed in the context of these new kinetic and mechanistic results.  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