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.
The enthalpies of formation (ΔH f o) for 23 halosubstituted radicals were determined from the published data on bond dissociation energies. The ΔH f o values of the corresponding molecules necessary for the calculation of ΔH f o of the radicals were taken from handbooks or calculated by the additive-group method. The conjugation energies of the radicals are calculated, and the effect of substituents at the π-system on these values was shown. Errors of determination of the ΔH f o values of the radicals were estimated. For Part 1, see Ref. 1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 643–646, April, 1998.  相似文献   

2.
The enthalpies of formation (ΔH f o) for 24 hydrocarbon radicals (R?), mainly polycyclic aromatic radicals with the complex structure, were determined from the published data on bond dissociation energies. The ΔH f o values of the corresponding molecules were calculated, in the majority of cases, by the macroincrement method. Calculations by the group contribution method were performed. Some ΔH f o(R?) values were compared to those calculated by the additive-group method. Calculations were performed, and the conjugation energies of the radicals were discussed. The errors of determination of the ΔH f o(R?) values found were estimated. Due to this work, the database for ΔH f o values of hydrocarbon radicals was increased more than by 25%.  相似文献   

3.
The formation enthalpies (H f°) of 12 organosulfur radicals (R·) were determined for the first time from the published values of dissociation energies of R—X bonds.  相似文献   

4.
The values of C−H and C−I bond dissociation energies were used to calculate the enthalpies of formation (δH f o of 20 cyclic and conjugated hydrocarbon radicals (R′). The values of δH f o (R′) were analyzed in terms of the quantitative structure-property correlation based on the additive-group model, and the reliability of these data was shown. Based on the correlation, several strain energies of cycles and energies of conjugation of a lone electron with a ρ-system were calculated. The additive-group method for calculation of δH f o can be extended for radicals of the naphthalyl type. For Part 2, see Ref. 1. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 286–288, February, 1999.  相似文献   

5.
[reaction: see text] Theoretical calculations of carbon-oxygen bond dissociation enthalpies in substituted methylperoxyl radicals (YCH(2)OO*) reveal that bond strengths are not governed solely by the stability of YCH2* radicals but are largely affected by hyperconjugation when Y is electron-donating or conjugating. In many cases, this hyperconjugative effect is greater than stabilization of the methyl radical by Y. All electron-withdrawing Y exert small destabilizing effects via inductive withdrawal of electrons from the polarized C-OO* bond.  相似文献   

6.
The enthalpies of formation (H f°) of 16 alkyl-substituted phenyl and benzyl radicals (R·) were determined for the first time by the published values of energies of R—X bond dissociation. For the initial molecules of RX, alkyl-substituted benzenes, the additive-group procedure was developed for the calculation of H f°. In the framework of the additive-group model for considered R·, we studied the structure-property interrelation, analyzed the obtained H f°(R·) values, and confirmed their reliability. The influence of nonvalent interactions on H f°(R·) was systematized and detailed. The parameters, from which it is possible to calculate H f° of the 51 radicals, were proposed.  相似文献   

7.
The balance between electrostatic and non-electrostatic enthalpic contributions to the free energy of solvation of a series of neutral solutes in water and n-octanol is examined by means of continuum solvation calculations based on the Miertus–Scrocco–Tomasi (MST) method. The experimental data indicate that the solvation enthalpy of hydrocarbons is very similar in water and n-octanol, and that the enthalpic contribution measured for polar compounds is larger in water than in n-octanol. According to MST calculations, the different magnitude of the solvation enthalpy found for polar compounds in the two solvents can be largely attributed to the electrostatic contribution. Moreover, the results point out that there is close resemblance between the non-electrostatic components for both hydrocarbons and polar compounds in the two solvents. Finally, the results show the power of current continuum models like MST to dissect the total free energy of solvation in entropic and enthalpic contributions and suggest that new refinements of continuum solvation models should include not only the fitting to solvation free energies, but also their enthalpic components.  相似文献   

8.
The bond dissociation enthalpies (BDEs) in chlorinated benzenes and phenols and the standard gas‐phase enthalpies of formation of chlorinated phenyl and phenoxy radicals are predicted by using Gaussian‐4 (G4) and Gaussian‐3X (G3X) model chemistries. The predicted G4 BDEs are systematically smaller than the G3X ones, with difference as much as ~15 kJ/mol for the C6Cl5‐Cl bond, and the G4 enthalpies of formation of the free radicals are systematically smaller than the G3X ones. The discrepancies increase gradually with the degree of chlorination; whereas for the closed‐shell species, G4 and G3X enthalpies of formation agree closely within 2 kJ/mol. The difference between G4 and G3X arises mainly from the noncanceling high‐level correction terms in G4. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 43: 62–69, 2011  相似文献   

9.
Theoretical calculations were carried out to provide a framework for understanding the free radical oxidation of unsaturated lipids. The carbon[bond]hydrogen bond dissociation enthalpies (BDEs) of organic model compounds and oxidizable lipids (R[bond]H) and the carbon[bond]oxygen bond dissociation enthalpies of peroxyl radical intermediates (R[bond]OO*) have been calculated. The carbon[bond]hydrogen BDEs correlate with the rate constant for propagation of free radical autoxidation, and the carbon[bond]oxygen BDEs of peroxyl radicals correlate with rate constants for beta-fragmentation of these intermediates. Oxygen addition to intermediate carbon radicals apparently occurs preferentially at centers having the highest spin density. The calculated spin distribution therefore provides guidance about the partitioning of oxygen to delocalized carbon radicals. Where the C[bond]H BDEs are a function of the extent of conjugation in the parent lipid and the stability of the carbon radical derived therefrom, C[bond]OO* BDEs are also affected by hyperconjugation. This gives way to different rates of beta-fragmentation of peroxyl radicals formed from oxygen addition at different sites along the same delocalized radical. We have also studied by both theory and experiment the propensity for benzylic radicals to undergo oxygen addition at their ortho and para carbons which, combined, possess an equivalent unpaired electron spin density as the benzylic position itself. We find that the intermediate peroxyl radicals in these cases have negative C[bond]OO* BDEs and, thus, have rate constants for beta-fragmentation that exceed the diffusion-controlled limit for the reaction of a carbon-centered radical with oxygen.  相似文献   

10.
Laccase is a multicopper oxidase that contains four Cu ions, one type 1, one type 2, and a coupled binuclear type 3 Cu pair. The type 2 and type 3 centers form a trinuclear Cu cluster that is the active site for O(2) reduction to H(2)O. To examine the reaction between the type 2/type 3 trinuclear cluster and dioxygen, the type 1 Cu was removed and replaced with Hg(2+), producing the T1Hg derivative. When reduced T1Hg laccase is reacted with dioxygen, a peroxide intermediate (P) is formed. The present study examines the kinetics and mechanism of formation and decay of P in T1HgLc. The formation of P was found to be independent of pH and did not involve a kinetic solvent isotope effect, indicating that no proton is involved in the rate-determining step of formation of P. Alternatively, pH and isotope studies on the decay of P revealed that a proton enhances the rate of decay by 10-fold at low pH. This process shows an inverse k(H)/k(D) kinetic solvent isotope effect and involves protonation of a nearby residue that assists in catalysis, rather than direct protonation of the peroxide. Decay of P also involves a significant oxygen isotope effect (k(16)O(2)/k(18)O(2)) of 1.11 +/- 0.05, indicating that reductive cleavage of the O-O bond is the rate-determining step in the decay of P. The activation energy for this process was found to be approximately 9.0 kcal/mol. The exceptionally slow rate of decay of P is explained by the fact that this process involves a 1e(-) reductive cleavage of the O-O bond and there is a large Franck-Condon barrier associated with this process. Alternatively, the 2e(-) reductive cleavage of the O-O bond has a much larger driving force which minimizes this barrier and accelerates the rate of this reaction by approximately 10(7) in the native enzyme. This large difference in rate for the 2e(-) versus 1e(-) process supports a molecular mechanism for multicopper oxidases in which O(2) is reduced to H(2)O in two 2e(-) steps.  相似文献   

11.
The O-O bond strengths in ten organic hydrotrioxides have been calculated by semiempirical MNDO and AMI methods. The RO-OOH bond strength is independent of the nature of substituent R and is equal to 20.4±1.1 kcal mol–1 (AM1). The influence of the inductive effect of substituent R on the value ofD(ROO-OH) has been established.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 1129–1131, May, 1996.  相似文献   

12.
Theoretical studies of O-O bond formation in photosystem II   总被引:1,自引:0,他引:1  
The most critical part of dioxygen evolution in photosystem II is the O-O bond formation step. In order to reach an efficient mechanism, nature uses a unique oxygen-evolving complex (OEC) having four manganese and one calcium center. Even though the structure of the OEC has become much more clear during recent years, it has still been difficult to find a transition state (TS) for O-O bond formation with a sufficiently low barrier. However, about a year ago, a quite surprising type of TS was found. With the latest X-ray ligand assignment, the local barrier for this TS is only 5.1 kcal/mol. It can be described as an attack by an oxygen radical, held by a dangling manganese, on a bridging oxo ligand in the Mn3Ca cube. In the present short Article, energy diagrams describing the entire process of dioxygen formation will be presented. An important conclusion drawn from these diagrams is that the major features of dioxygen formation remain the same irrespective of which one of the experimentally suggested structures the diagram is built on. Compared to earlier presentations of the same type, a slightly different approach has been used for setting up the diagrams. Results from a recent experimental study of the pressure dependence of oxygen release have been used to define the final energy levels. The loss of energy in the electron transfer from Tyrz to P680 has also been incorporated into the diagrams. A good agreement with experimental observations is demonstrated.  相似文献   

13.
Low-temperature oxygenation of copper(I) complexes of N,N,N',N'-tetraethylpropane-1,3-diamine yields solutions containing both mu-eta2:eta2-peroxodicopper(II) (P) and bis(mu-oxo)dicopper(III) (O) valence isomers. The P/O equilibrium position depends on the nature of the counteranion; P is favored with more basic anions. Titration and EXAFS experiments as well as DFT calculations suggest that axial donation from a sulfonate anion to the copper centers imparts an electronic/electrostatic bias toward the P isomer.  相似文献   

14.
The O-H bond dissociation enthalpies (BDEs) of 13 oximes, RR'C=NOH, having R and/or R' = H, alkyl, and aryl are reported. Experimental anchor points used to validate the results of theoretical calculations include (1) the O-H BDEs of (t-Bu)2C=NOH, t-Bu(i-Pr)C=NOH, and t-Bu(1-Ad)C=NOH determined earlier from the heat released in the reaction of (t-Bu)2C=NO* with (PhNH)2 in benzene and EPR spectroscopy (Mahoney, L. R.; Mendenhall, G. D.; Ingold, K. U. J. Am. Chem. Soc. 1973, 95, 8610), all of which were decreased by 1.7 kcal/mol to reflect a revision to the heat of formation of (E)-azobenzene (which has significant ramifications for other BDEs) and to correct for the heat of hydrogen bonding of (t-Bu)2C=NOH (alphaH2 = 0.43 measured in this work) to benzene, and (2) the measured rates of thermal decomposition of six RR'C=NOCH2Ph at 423 or 443 K, which were used to derive O-H BDEs for the corresponding RR'C=NOH. Claims (Bordwell, F. G.; Ji, G. Z. J. Org. Chem. 1992, 57, 3019; Bordwell, F. G.; Zhang, S. J. Am. Chem. Soc. 1995, 117, 4858; and Bordwell, F. G.; Liu, W.-Z. J. Am. Chem. Soc. 1996, 118, 10819) that the O-H BDEs in mono- and diaryloximes are significantly lower than those for alkyloximes due to delocalization of the unpaired electron into the aromatic ring have always been inconsistent with the known structures of iminoxyl radicals as are the purported perpendicular structures, i.e., phi(Calpha-C=N-O*) = 90 degrees, for sterically hindered dialkyl iminoxyl radicals. The present results confirm the 1973 conclusion that simple steric effects, not electron delocalization or dramatic geometric changes, are responsible for the rather small differences in oxime O-H BDEs.  相似文献   

15.
A detailed reactivity study has been carried out for the first time on a new mononuclear alkylperoxo copper(II) complex, which is generated by the reaction of copper(II) complex supported by the bis(pyridylmethyl)amine tridentate ligand containing a phenyl group at the 6-position of the pyridine donor groups and cumene hydroperoxide (CmOOH) in CH3CN. The cumylperoxo copper(II) complex thus obtained has been found to undergo homolytic cleavage of the O-O bond and induce C-H bond activation of exogenous substrates, providing important insights into the catalytic mechanism of copper monooxygenases.  相似文献   

16.
17.
Enthalpies of formation (Δ f H o) of 57 oxygen-containing heterocyclic radicals have been determined from literature values of the dissociation energies of C–H chemical bonds in appropriate molecules. Within the framework of a quantitative structure–property correlation based on the additive-group model, an analysis has been carried out of the obtained Δ f H o values, the reliability of these data has been shown, and a set of values has been determined of the cyclic stress of the radicals being considered. The parameters of the additive-group method are recommended for the calculation of Δ f H o for radicals of the considered class.  相似文献   

18.
A computational study on the thermodynamic properties of 13 beta-diketones is presented. The B3LYP//6-311+G(2d,2p)//B3LYP/6-31G(d) theoretical approach was employed to compute the O-H and C-H bond dissociation enthalpies and enthalpy of tautomerization and to estimate standard gas-phase enthalpies of formation for the radicals and for the parent molecules. The gas-phase enthalpies of formation for the neutral molecules are in excellent agreement with available experimental data, supporting the estimates made for the radicals. The latter are very important for the clarification of the thermochemistry of many beta-diketonato metal complexes previously reported in the literature. Importantly, when substituents R = -CHR' are attached to the beta-diketone's scaffold, C-H homolytic bond cleavage is always favored with respect to O-H bond scission.  相似文献   

19.
Peroxynitrite decay in weakly alkaline media occurs by two concurrent sets of pathways which are distinguished by their reaction products. One set leads to net isomerization to NO(3)(-) and the other set to net decomposition to O(2) plus NO(2)(-). At sufficiently high peroxynitrite concentrations, the decay half-time becomes concentration-independent and approaches a limiting value predicted by a mechanism in which reaction is initiated by unimolecular homolysis of the peroxo O-O bond, i.e., the following reaction: ONOOH --> (*)OH + (*)NO(2). This dynamical behavior excludes alternative postulated mechanisms that ascribe decomposition to bond rearrangement within bimolecular adducts. Nitrate and nitrite product distributions measured at very low peroxynitrite concentrations also correspond to predictions of the homolysis model, contrary to a recent report from another laboratory. Additionally, (1) the rate constant for the reaction ONOO(-) --> (*)NO + (*)O(2)(-), which is critical to the kinetic model, has been confirmed, (2) the apparent volume of activation for ONOOH decay (DeltaV() = 9.7 +/- 1.4 cm(3)/mol) has been shown to be independent of the concentration of added nitrite and identical to most other reported values, and (3) complex patterns of inhibition of O(2) formation by radical scavengers, which are impossible to rationalize by alternative proposed reaction schemes, are shown to be quantitatively in accord with the homolysis model. These observations resolve major disputes over experimental data existing in the literature; despite extensive investigation of these reactions, no verifiable experimental evidence has been advanced that contradicts the homolysis model.  相似文献   

20.
Density functional theory (B3LYP) has been applied to large models of the Fe(II)-Cu(I) binuclear center in cytochrome oxidase, investigating the mechanism of O-O bond cleavage in the mixed valence form of the enzyme. To comply with experimental information, the O(2) molecule is assumed to be bridging between iron and copper during the O-O bond cleavage, leading to the formation of a ferryl-oxo group and a cupric hydroxide. In accord with previous suggestions, the calculations show that it is energetically feasible to take the fourth electron needed in this reaction from the tyrosine residue that is cross-linked to one of the copper ligands, resulting in the formation of a neutral tyrosyl radical. However, the calculations indicate that simultaneous transfer of an electron and a proton from the tyrosine to dioxygen during bond cleavage leads to a barrier more than 10 kcal/mol higher than that experimentally determined. This may be overcome in two ways. If an extra proton in the binuclear center assists in the mechanism, the calculated reaction barrier agrees with experiment. Alternatively, the fourth electron might initially be supplied by a residue in the vicinity other than the tyrosine.  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