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.
Theoretical calculations of ESR parameters for aminoxyl radicals have been widely studied using the density functional theory (DFT) calculations. However, the isomer N‐alkoxyaminyl radicals have been limitedly studied. With the use of experimental data for 46 N‐alkoxyaminyl and 38 aminoxyl radicals, the isotropic 14N hyperfine coupling constants (aN) and g‐factors have been theoretically estimated by several DFT calculations. The best calculation scheme of aN for N‐alkoxyaminyl radicals was PCM/B3LYP/6‐31 + + G(d,p) (R2 = 0.9519, MAE = 0.034 mT), and that for aminoxyl radicals was PCM/BHandHLYP/6‐31 + + G(3df,3pd) (R2 = 0.9336, MAE = 0.057 mT). For aminoxyl radicals, the solvation models in calculations enhanced the accuracy of reproducibility. In contrast, for N‐alkoxyaminyl radicals the calculations with solvation models provided no improvement. The differences in the best functionals between two types of radicals were thought to come from the contribution ratios of neutral and dipolar canonical structures in resonance forms. The aN for N‐alkoxyaminyl radicals that were stabilized by small contribution of dipolar canonical structures could be precisely reproduced by B3LYP with only 20% HF exact exchange. In contrast, the aN for aminoxyl radicals stabilized by large contribution of dipolar canonical structures was well reproduced by BHandHLYP with 50% HF exchange. The best calculation scheme of g‐factors was IEFPCM/B3LYP/6‐31 + G(d,p) (R2 = 0.9767, MAE = 0.0001) for not only aminoxyl but also N‐alkoxyaminyl radicals. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

2.
Quantum chemical calculations (DFT/B3LYP/6-31G(d)) considering specific solvation effects were used to compare the thermodynamic parameters of electrophilic substitution reactions (with the hydroxonium ion as a model electrophile) in 1H-tetrazole according to the addition—elimination and elimination—addition schemes. The latter scheme can proceed without preliminary formation of N-protonated azolium salts, as demonstrated earlier by the DFT/ B3LYP/6-31G(d,p) and DFT/B3LYP/6-31G(2df,p) calculations considering the solvation effects in aqueous solution in terms of the polarizable continuum model (PCM) with a proton as a model electrophile.  相似文献   

3.
The geometries and electronic properties of substrates, transition structures (TS), and product radicals in modeled elementary propagation reactions were studied for the styrene–acrylonitrile monomer system by use of quantum‐mechanical calculations: (DFT/B3‐LYP/6–31G(d), ROMP2/6–311+G(3df,2p)//DFT/B3‐LYP/6–31G(d), and DFT/B3‐LYP/6–311+G(3df,2p)//DFT/B3‐LYP/6–31G(d)) and for some parameters, the high‐level composite method G3 (Gaussian‐3, G3/MP2). Activation enthalpies (ΔHact) and reaction enthalpies (ΔHr) for modeled propagation reactions at 298.15 K were evaluated. The enthalpy of activation energy (ΔHact, kJ/mol) for the investigated elementary reactions rises for the B3‐LYP calculation in the following order: (CH3A?+S) < (CH3A?+A) < (CH3S?+A) < (CH3S?+S). For three propagation reactions, (CH3A?+A), (CH3A?+S), and (CH3S?+A), correlation between reaction enthalpy and enthalpy of activation suggests weak or negligible polar effects reflecting the Evans–Polanyi relation. However, from the electron affinities and ionization energies values data, it is not excluded that at least for [CH3A?+S[b]] and [CH3S?+A[b]] reactions, nucleophilic and electrophilic polar effects, respectively, can also be expected. The dependencies between TS geometries, electronic parameters, and enthalpic effects suggest the presence of a steric factor in all TS, including its exceptionally high contribution to the activation enthalpy for the CH3S?+S addition. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 1827–1844, 2005  相似文献   

4.
Intramolecular H‐atom transfer in model peptide‐type radicals was investigated with high‐level quantum‐chemistry calculations. Examination of 1,2‐, 1,3‐, 1,5‐, and 1,6[C ? N]‐H shifts, 1,4‐ and 1,7[C ? C]‐H shifts, and 1,4[N ? N]‐H shifts (Scheme 1), was carried out with a number of theoretical methods. In the first place, the performance of UB3‐LYP (with the 6‐31G(d), 6‐31G(2df,p), and 6‐311+G(d,p) basis sets) and UMP2 (with the 6‐31G(d) basis set) was assessed for the determination of radical geometries. We found that there is only a small basis‐set dependence for the UB3‐LYP structures, and geometries optimized with UB3‐LYP/6‐31G(d) are generally sufficient for use in conjunction with high‐level composite methods in the determination of improved H‐transfer thermochemistry. Methods assessed in this regard include the high‐level composite methods, G3(MP2)‐RAD, CBS‐QB3, and G3//B3‐LYP, as well as the density‐functional methods B3‐LYP, MPWB1K, and BMK in association with the 6‐31+G(d,p) and 6‐311++G(3df,3pd) basis sets. The high‐level methods give results that are close to one another, while the recently developed functionals MPWB1K and BMK provide cost‐effective alternatives. For the systems considered, the transformation of an N‐centered radical to a C‐centered radical is always exothermic (by 25 kJ ? mol?1 or more), and this can lead to quite modest barrier heights of less than 60 kJ ? mol?1 (specifically for 1,5[C ? N]‐H and 1,6[C ? N]‐H shifts). H‐Migration barriers appear to decrease as the ring size in the transition structure (TS) increases, with a lowering of the barrier being found, for example when moving from a rearrangement proceeding via a four‐membered‐ring TS (e.g., the 1,3[C ? N]‐H shift, CH3? C(O)? NH..CH2? C(O)? NH2) to a rearrangement proceeding via a six‐membered‐ring TS (e.g., the 1,5[C ? N]‐H shift, .NH? CH2? C(O)? NH? CH3 → NH2? CH2? C(O)? NH? CH2.).  相似文献   

5.
OH addition reactions play a pivotal role in the atmospheric transformation of a number of phenyl and substituted phenyl‐based persistent and toxic organic pollutants. Here, we screened appropriate DFT functionals to predict reaction mechanisms and rate constants (kOH) of the OH additions by taking benzene and substituted benzenes (C6H5F, C6H5Cl, C6H5Br, C6H5CH3, C6H5OH) as model compounds. By comparing the kOH values calculated with DFT methods to experimental values, we found that the ωB97 functional is the best among the 18 functionals considered (using the basis sets 6‐31 + G(d,p) for optimizations and 6‐311++G(3df,2pd) for single point energy calculations) in the temperature range of 230‐330 K. In addition, we found that some other functionals performed well in specific conditions, e.g., BMKD3 is good for benzene, halogenated benzenes and C6H5CH3, and CAM‐B3LYP is good for the reaction of C6H5OH at room temperature. Based on the diversity of the electronic structures of the selected model compounds and the frequent occurrence of certain substituents ( CH3,  OH,  F,  Cl, and  Br) in the target compounds, the functionals recommended here can be used for future study of the reaction mechanisms and kOH values for OH addition to phenyl and substituted phenyl‐based persistent and toxic organic pollutants.  相似文献   

6.
The potential energy profiles for the mutual conversion of the isomeric molecular ions [C5H6O]+? of 2‐methylfuran, 3‐methylfuran and 4H‐pyran and the fragmentations that lead to [C5H5O]+ ions were obtained from calculations at the B3LYP/6‐311G + + (3df,3pd)//B3LYP/6‐31G(d,p) level of theory. The various competing unimolecular processes were characterized by their RRKM microcanonical rate coefficients, k(E), using the sets of reactant and transition state frequencies and the kinetic barriers obtained from the density functional method. In either a high‐ or a low‐energy regime, the pyrylium ion [C5H5O]+ is generated directly from the 4H‐pyran molecular ion by a simple cleavage. In contrast, in the metastable kinetic window, the molecular ions of methylfurans irreversibly isomerize to a mixture of interconverting structures before dissociation, which includes the 2H‐ and 3H‐pyran ions. The hydrogen atoms attached to saturated carbons of the pyran rings are very stabilizing at the position 2, but they are very labile at position 3 and can be shifted to adjacent positions. Once 4H‐pyran ion has been formed, the C? H bond cleavage begins before any hydrogen shift occurs. According to our calculation, there would not be complete H scrambling preceding the dissociation of the molecular ions [C5H6O]+?. On the other hand, as the internal energy of the 2‐methylfuran molecular ion increases, H? loss can become more important. These results agree with the available experimental data. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

7.
Below −60° and without catalyst, 1,2‐dimethylidenecyclopentane ( 16 ), 1,2‐dimethylidenecyclohexane ( 13 ), 1,2‐dimethylidenecycloheptane ( 17 ), and 1,2‐dimethylidenecyclooctane ( 18 ) add to sulfur dioxide in the hetero‐Diels‐Alder mode, giving the corresponding sultines 4,5,6,7‐tetrahydro‐1H‐cyclopent[d][1,2]oxathiin 3‐oxide ( 19 ), 1,4,5,6,7,8‐hexahydro‐2,3‐benzoxathiin 3‐oxide ( 14 ), 4,5,6,7,8,9‐hexahydro‐1H‐cyclohept[d][1,2]oxathiin 3‐oxide ( 20 ), and 1,4,5,6,7,8,9,10‐octahydrocyclooct[d][1,2]oxathiin 3‐oxide ( 21 ), respectively. Above −40°, the sultines are isomerized into the corresponding sulfolenes 3,4,5,6‐tetrahydro‐1H‐cyclopenta[c]thiophene 2,2‐dioxide ( 22 ), 1,3,4,5,6,7‐hexahydrobenzo[c]thiophene 2,2‐dioxide ( 15 ), 3,4,5,6,7,8‐hexahydro‐1H‐cyclohepta[c]thiophene 2,2‐dioxide ( 23 ), and 1,3,4,5,6,7,8,9‐octahydrocycloocta[c]thiophene 2,2‐dioxide ( 24 ). Kinetics and thermodynamics data were collected for these reactions. The sultines are ca. 10 kcal/mol Diels‐Alder additions (ΔH( 16 −36±3 cal mol−1 K−1) in agreement with third‐order rate laws that imply that two molecules of SO2 intervene in the transition states of these cycloadditions. Similar observations were made for the cheletropic additions of SO2. Attempts to simulate the thermodynamics and kinetics parameters of the reactions of SO2 with dienes 16 and 13 by density‐functional theory (DFT) suggest that the calculations require an appropriate number of polarization functions in the basis set employed. A satisfactory recipe to compute the SO2 additions to large dienes can be: B3LYP/6‐31G(d) geometry optimizations followed by B3LYP/6‐31+G(2df,p) single‐point calculations or G2(MP2,SVP) estimates on the B3LYP/6‐31G(d) geometries.  相似文献   

8.
杨颙  张为俊  高晓明 《中国化学》2006,24(7):887-893
A theoretical study on the blue-shifted H-bond N-H…O and red-shifted H-bond O-H…O in the complexHNO…H_2O_2 was conducted by employment of both standard and counterpoise-corrected methods to calculate thegeometric structures and vibrational frequencies at the MP2/6-31G(d),MP2/6-31 G(d,p),MP2/6-311 q G(d,p),B3LYP/6-31G(d),B3LYP/6-31 G(d,p) and B3LYP/6-311 G(d,p) levels.In the H-bond N-H…O,the calcu-lated blue shift of N-H stretching frequency is in the vicinity of 120 cm~(-1) and this is indeed the largest theoreticalestimate of a blue shift in the X-H…Y H-bond ever reported in the literature.From the natural bond orbital analy-sis,the red-shifted H-bond O-H…O can be explained on the basis of the dominant role of the hyperconjugation.For the blue-shifted H-bond N-H…O,the hyperconjugation was inhibited due to the existence of significant elec-tron density redistribution effect,and the large blue shift of the N-H stretching frequency was prominently due tothe rehybridization of sp~n N-H hybrid orbital.  相似文献   

9.
This work puts forth a reaction pathway for the reactivation of exogenous ligand inhibited H‐cluster, the active site of Fe‐only hydrogenases. The H‐cluster is a dimetal complex, Fe–Fe, with the metal centers bridged by di(thiomethyl)amine. Exogenous ligands, H2O, and OH?, are bound to the distal iron (Fed). Density functional theory (DFT) calculations on the native and ruthenium‐modified H‐cluster have been performed using the B3LYP functional with 6‐31+G** and 6‐311+G** basis sets. We have ascertained that there is a thermodynamically favorable pathway for the reactivation of the OH? inhibited H‐cluster, which proceeds by an initial protonation of the Fed–OH? complex. The proposed reaction pathway has all its intermediate reactions ensue exothermically. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

10.
On the basis of DFT calculations (B3LYP/6‐311+G**), the possibility to include solvent effects is considered in the investigation of the H2O‐exchange mechanism on [Be(H2O)4]2+ within the widely used cluster approach. The smallest system in the gas phase, [Be(H2O)4(H2O)]2+, shows the highest activation barrier of +15.6 kcal/mol, whereas the explicit addition of five H‐bonded H2O molecules in [{Be(H2O)4(H2O)}(H2O)5]2+ reduces the barrier to +13.5 kcal/mol. Single‐point calculations applying CPCM (B3LYP(CPCM:H2O)/6‐311+G**//B3LYP/6‐311+G**) on [Be(H2O)4(H2O)]2+ lower the barrier to +9.6 kcal/mol. Optimization of the precursor and transition state of [Be(H2O)4(H2O)]2+ within an implicit model (B3LYP(CPCM:H2O)/6‐311+G** or B3LYP(PCM:H2O)/6‐311+G**) reduces the activation energy further to +8.3 kcal/mol but does not lead to any local minimum for the precursor and is, therefore, unfavorable.  相似文献   

11.
In this study density functional theory (DFT) calculations at B3LYP/6-31G(d), B3LYP/6-31+G(d) and B3LYP/6-311+G(2df,2p) levels for geometry optimization and total energy calculation were applied for investigation of the important energy-minimum conformations and transition-state of 1,2-, 1,3-, and 1,4-dithiepanes. Moreover, ab initio calculations at HF/6-31G(d) level of theory for geometry optimization and MP2/6-311G(d)//HF/ 6-31G(d) level for a single-point total energy calculation were reported for different conformers. The obtained results reveal that, the twist-chair conformer is a global minimum for all of these compounds. Also, two local minimum were found in each case, which are twisted-chair and twisted-boat conformers. The boat and chair geometries are transition states. The minimum energy conformation of 1,2-dithiepane is more stable than the lowest energy forms of 1,3-dithiepane and 1,4-dithiepane. Furthermore, the anomeric effect was investigated for 1,3-dithiepane by the natural bond orbital method. The computational results of this study shows that all conformers of 1,3-dithiepane have a hypercojugation system. Finally, the 13C NMR chemical shifts for the conformers of 1,4-dithiepane were calculated, which have good correlation with their experimental values.  相似文献   

12.
We report a study on different ionization states and conformations of the bimolecular (Gly)2 system by means of quantum mechanical calculations. Optimized geometries for energy minima of the glycine dimer, as well as relative energies and free energies were computed as a function of the medium: gas phase, nonpolar polarizable solvent, and aqueous solution. The polarizable continuum model was employed to account for solvation effects. Energy calculations were done using the MP2/aug‐cc‐pVTZ and B3LYP/6‐311+G(2df,2p) methods on B3LYP/6‐31+G(d,p) optimized structures (some single‐point energy calculations were also done using the B3PW91 and PBE1KCIS methods). Ionized forms of the glycine dimer (either zwitterion–zwitterion or neutral–zwitterion) are predicted to exist in all media, in contrast to amino acid monomers. In aqueous solution, dimerization is an exergonic process (?4 kcal mol?1). Thus, according to our results, zwitterion–zwitterion Gly dimers might be abundant in supersaturated glycine aqueous solutions, a fact that has been connected with the structure of α‐glycine crystals but that remains controversial in the literature. Another noticeable result is that zwitterion–zwitterion interactions are substantially underestimated when computed using methods based on density functional theory. For comparison, some calculations for the dimer of the simplest chiral amino acid alanine were done as well and differences to the glycine dimer are discussed.  相似文献   

13.
The hydrolytic deamination reaction mechanism of guanine (G) and (H2O)n (n = 1–4) has been theoretically investigated, including solvent effects at the B3LYP/6-31G (d,p) and MP2/6-311G (d,p) levels. The results show that the hydrolytic deamination reaction of G involves two steps. In the first step, a tetra-coordinated intermediate is generated by a hydrolysis reaction, and then deamination reaction is carried out. In the second step, the final products form via a hydrogen transfer. For the hydrolytic deamination reaction of guanine and 3H2O or 4H2O, only two molecules of water participate in the reaction, and other water molecule make the transition state more stable as a catalyst. The study on the potential energy surface shows that the deamination reaction of G and nH2O does not take place because of a higher barrier for the opening system, which is in agreement with the experimental result.  相似文献   

14.
The title molecular salt, 4-(2-hydroxyphenyl)-4,5,6,7-tetrahydro-1H-imidazo[4,5-c]pyridin-5-ium chloride hydrate (C12H14N3O+·Clˉ·H2O), was synthesized and characterized by IR-NMR spectroscopy and single-crystal X-ray diffraction. In addition to the molecular geometry from X-ray experiment, the molecular geometry, vibrational frequencies and gauge-independent atomic orbital (GIAO) 1H and 13C NMR chemical shift values of the title compound in the ground state have been calculated using the density functional theory (DFT/B3LYP) method with the 6-31++G(d,p) and 6-311++G(d,p) basis sets, and compared with the experimental data. Besides, molecular electrostatic potential (MEP) distribution and non-linear optical properties of the title compound were investigated by theoretical calculations at the B3LYP/6-311++G(d,p) level.  相似文献   

15.
The title molecule, 3‐{[4‐(3‐methyl‐3‐phenyl‐cyclobutyl)‐thiazol‐2‐yl]‐hydrazono}‐1,3‐dihydro‐indol‐2‐one (C22H20N4O1S1), was prepared and characterized by 1H NMR, 13C NMR, IR, UV–visible, and single‐crystal X‐ray diffraction. The compound crystallizes in the monoclinic space group P21 with a = 8.3401(5), b = 5.6976(3), c = 20.8155(14) Å, and β = 95.144(5)°. Molecular geometry from X‐ray experiment and vibrational frequencies of the title compound in the ground state has been calculated using the Hartree–Fock with 6‐31G(d, p) and density functional method (B3LYP) with 6‐31G(d, p) and 6‐311G(d, p) basis sets, and compared with the experimental data. The calculated results show that optimized geometries can well reproduce the crystal structural parameters, and the theoretical vibrational frequencies values show good agreement with experimental data. Density functional theory calculations of the title compound and thermodynamic properties were performed at B3LYP/6‐31G(d, p) level of theory. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012  相似文献   

16.
Heteroaromatic hydrocarbons (including thiophene [TH], benzothiophene [BT], and dibenzothiophene [DBT]) do not have apparent functional groups capable of interacting with the silica‐oxide tetrahedral surface of kaolinite. Thus, question remains concerning what would be the driving forces for the adsorption. Here, the Si13O37H22 cluster model for the surface is constructed, and the interactions of the surface with three heteroaromatic compounds are studied at the MP2/6‐31G(d,p)//B3LYP/6‐31G(d) level. The computed properties characterizing the complexes include optimized structural parameters, electron density characteristics (the ρ and ? 2ρ values for C? H…O bonds), adsorption energies, vibration frequencies and electrostatic potential maps. The results suggest that the C? H…O hydrogen bonding interactions between the heteroaromatic compounds and tetrahedral surface are likely among the important interactions for the adsorption. The order of the stability of the cluster model of kaolinite complexed with the heteroaromatic compounds is found to be 3Si? O? DBT > 3Si? O? BT > 3Si? O? TH based on the calculations.  相似文献   

17.
The interaction of bare iron mono‐ and dications with hydrogen peroxide in the gas phase is studied by ab initio calculations employing the B3LYP/6‐311+G* level of theory. For the monocation, the quartet and sextet coordination complexes Fe(H2O2) are high‐energy isomers that easily interconvert to the more stable iron dihydroxide monocation Fe(OH) and hydrated iron oxide (H2O)FeO+ (quartet) or dissociate into FeOH++OH. (sextet). On the dication surface, however, the order of stabilities is reversed in that Fe(H2O2)2+ (quintet) corresponds to the most stable doubly charged species, while the formal FeIV compounds Fe(OH) and (H2O)FeO2+ are higher in energy.  相似文献   

18.
Density functional theory (DFT) based calculations are performed on a series of alkyl nitrites and nitroalkanes representing large‐scale primary, secondary, and tertiary nitro compounds and their radicals resulting from the loss of their skeletal hydrogen atoms. Geometries, vibration frequencies, and thermochemical properties [S°(T) and C°p(T) (10 K ? T ? 5000 K)] are calculated at the B3LYP/6‐31G(d,p) DFT level. Δf298 values are from B3LYP/6‐31G(d,p), B3LYP/6‐31+G(2d,2p), and the composite CBS‐QB3 levels. Potential energy barriers for the internal rotations have been computed at the B3LYP/6‐31G(d,p) level of theory, and the lower barrier contributions are incorporated into entropy and heat capacity data. The standard enthalpies of formation at 298 K are evaluated using isodesmic reaction schemes with several work reactions for each species. Recommended values derived from the most stable conformers of respective nitro‐ and nitrite isomers include ?30.57 and ?28.44 kcal mol?1 for n‐propane‐, ?33.89 and ?32.32 kcal mol?1 for iso‐propane‐, ?42.78 and ?41.36 kcal mol?1 for tert‐butane‐nitro compounds and nitrites, respectively. Entropy and heat capacity values are also reported for the lower homologues: nitromethane, nitroethane, and corresponding nitrites. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 181–199, 2010  相似文献   

19.
According to B3LYP/6-31G(d,p) and B3LYP/6-31G(d) calculations, intramolecular migrations of H+ and Na+ cations in the molecule of 7-ethyl-2,3,5,6,8-pentahydroxy-1,4-naphthoquinone (echinochrome A) monosodium salt in the gas phase and in clusters containing five water molecules can proceed asynchronously through the chain of O-H...O hydrogen bonds by the jump-over mechanism.  相似文献   

20.
We report herein the synthesis and characterization of a new proton sponge derivative, 1,8‐bis(bis(diisopropylamino)cyclopropeniminyl)naphthalene 4 (DACN), as well as its bis‐protonated counterpart 6 . A crystal structure of 6 is presented, along with variable temperature 1H NMR data on the BF4? salt ( 6?BF4 ). DFT calculations were performed to investigate the structure of the monoprotonated species 7 and to gain insight into the structural and electronic nature of all three species. The proton affinity (PA) of 4 , calculated at the B3LYP/6‐311G++(d,p)//B3LYP/6‐31G(d,p) level, taking into account thermal corrections from the B3LYP/6‐31G(d,p) method, was 282.3 kcal mol?1, while its pKa was estimated at 27.0. NICS calculations were performed to examine the changes in aromaticity within these systems upon each successive protonation. Lastly, homodesmotic reaction schemes were used in order to estimate the factors contributing to the strong PA predicted for 4 .  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