共查询到20条相似文献,搜索用时 15 毫秒
1.
Kelson C. Lopes Mozart N. Ramos Silmar A. do Monte Elizete Ventura Regiane C.M.U. Araú jo 《Journal of Molecular Structure》2006,760(1-3):21-27
The two-level factorial design (FD) and principal component analysis (PCA) chemometric techniques were used to investigate the carbon–hydrogen bond lengths dependence on the basis set size and quantum chemistry method, for H–C≡CH, H–C≡CF, H–C≡CCH3, H–C≡CCN, H–C≡CCl and H–C≡CCCH molecular systems. The calculations were performed by using Hartree-Fock (HF), Møller-Plesset 2 (MP2) and Density Functional Theory (DFT) with B3LYP exchange-correlation functional methods. The effects concerning basis set size include the number of valence and polarization functions as well as the cooperative effect between them, at all computational levels. The increase in the number of valence functions decreases the calculated C–H bond lengths by approximately 0.0022 Å, while the inclusion of polarization functions at HF and B3LYP levels increases the C–H bond length, in contrast to the behavior obtained at MP2 level. The effect of the inclusion of diffuse functions is non-significant, at all three computational levels. Moreover, the valence–polarization interaction effects are not significant, except at the MP2 calculational level, in which such effects lead to an increase in the calculated C–H bond lengths. When the computational level changes from HF→B3LYP and B3LYP→MP2 the calculated C–H bond length values increase (on average) by +0.0100 and +0.0027 Å, respectively. Algebraic models (one for each level of calculation) successfully employed to reproduce the calculated values for H–C≡N bond length, a system not included in the training set. The HF/6-31G(d,p) and HF/6-31++G(d,p) results yield the lowest standard errors (0.0015 and 0.0014 Å, respectively) and correspond to the calculated points in closest proximity to the experimental one. 相似文献
2.
Eudes E. Fileti Puspitapallab Chaudhuri Sylvio Canuto 《Chemical physics letters》2004,400(4-6):494-499
Hydrogen binding energies are calculated for the different isomers of 1:1 complexes of methanol, ethanol and water using ab initio methods from MP2 to CCSD(T). Zero-point energy vibration and counterpoise corrections are considered and electron correlation effects are analyzed. In methanol–water and ethanol–water the most stable heterodimer is the one where the water plays the role of proton donor. In methanol–ethanol the two isomers have essentially the same energy and no favorite heterodimer could be discerned. The interplay between the relative binding energy is briefly discussed in conjunction with the incomplete mixing of alcohol–water systems. 相似文献
3.
Gregory K. Pierens Taracad K. Venkatachalam David C. Reutens 《Magnetic resonance in chemistry : MRC》2016,54(12):941-946
Two‐ and three‐bond coupling constants (2JHC and 3JHC) were determined for a series of 12 substituted cinnamic acids using a selective 2D inphase/antiphase (IPAP)‐single quantum multiple bond correlation (HSQMBC) and 1D proton coupled 13C NMR experiments. The coupling constants from two methods were compared and found to give very similar values. The results showed coupling constant values ranging from 1.7 to 9.7 Hz and 1.0 to 9.6 Hz for the IPAP‐HSQMBC and the direct 13C NMR experiments, respectively. The experimental values of the coupling constants were compared with discrete density functional theory (DFT) calculated values and were found to be in good agreement for the 3JHC. However, the DFT method under estimated the 2JHC coupling constants. Knowing the limitations of the measurement and calculation of these multibond coupling constants will add confidence to the assignment of conformation or stereochemical aspects of complex molecules like natural products. Copyright © 2016 John Wiley & Sons, Ltd. 相似文献
4.
Ytterbium and samarium metals reduced aromatic ketimines to give directly divalent azalanthanacyclopropane complexes 1 quantitatively, the structure of which was characterized by X-ray analysis. The imine complexes 1 catalyzed dehydrogenative silylation of terminal alkynes, hydrosilylation of imines and alkenes, and intermolecular hydrophosphination of alkynes. Moreover, dehydrogenative double silylation of conjugated dienes was achieved with 1. 相似文献
5.
The dual-level direct dynamics approach is employed to study the dynamics of the CH(3)OCH(3) + H (R1) and CH(3)OCH(3) + CH(3) (R2) reactions. Low-level calculations of the potential energy surface are carried out at the MP2/6-311+G(d,p) level of theory. High-level energetic information is obtained at the QCISD(T) level of theory with the 6-311+G(3df,3pd) basis set. The dynamics calculations are performed using variational transition state theory (VTST) with the interpolated single-point energies (ISPE) method, and small-curvature tunneling (SCT) is included. It is shown that the reaction of CH(3)OCH(3) with H (R1) may proceed much easier and with a lower barrier height than the reaction with CH(3) radical (R2). The calculated rate constants and activation energies are in good agreement with the experimental values. The calculated rate constants are fitted to k(R1) = 1.16 x 10(-19) T(3) exp(-1922/T) and k(R2) = 1.66 x 10(-28) T(5) exp(-3086/T) cm(3) mol(-1) s(-1) over a temperature range 207-2100 K. Furthermore, a small variational effect and large tunneling effect in the lower temperature range are found for the two reactions. 相似文献
6.
The gas phase reaction I2 + HCOOCH3 → HI + CH3I + CO2 has been studied spectrophotometrically in a static system over the pressure ranges I2 (6–39 torr) and HCOOMe (28–360 torr). In the temperature range 293–356°, the initial rate of disappearance of I2 is first order in [HCOOMe] and half-order in [I2]. The rate determining step is where k1 is given by where θ = 2.303 RT in kcal/mole. This activation energy gives a carbonyl C? H bond strength of 92.7 kcal/mole. At 356° there was no evidence of abstraction of a methoxy hydrogen, so a lower limit of 100 kcal/mole may be placed on this C? H bond strength. These ester C? H bond strengths are discussed in relation to comparable values in aldehydes and ethers. 相似文献
7.
Methyl radical complexes H3C…HCN and H3C…HNC have been investigated at the UMP2(full)/aug‐cc‐pVTZ level to elucidate the nature of hydrogen bonds. To better understand the intermolecular H‐bond interactions, topological analysis of electron density at bond critical points (BCP) is executed using Bader's atoms‐in‐molecules (AIM) theory. Natural bond orbital (NBO) analysis has also been performed to study the orbital interactions and change of hybridization. Theoretical calculations show that there is no essential difference between the blue‐shift H‐bond and the conventional one. In H3C…HNC complex, rehybridization is responsible for shortening of the N? H bond. The hyperconjugative interaction between the single electron of the methyl radical and N? H antibonding orbital is up to 7.0 kcal/mol, exceeding 3.0 kcal/mol, the upper limit of hyperconjugative n(Y)→σ*(X–H) interaction to form the blue‐shifted H‐bond according to Alabugin's theory. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2007 相似文献
8.
Fluorous chiral BINOLs and BINAP were prepared and used as the ligands for an asymmetric addition of Et2Zn to aromatic aldehydes and an asymmetric Heck reaction, respectively. The enantioselectivities were similar in homogeneous system to those of the original non-fluorous reactions. Consecutive reactions were examined by utilizing fluorous–organic biphase and fluorous solid phase extraction techniques. Enantioselectivities in consecutive reactions were close to that attained in the non-fluorous system. The solid phase extraction method also enabled us to perform a simultaneous screening procedure. 相似文献
9.
《Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy》1982,38(1):113-118
Infrared spectra in the CH stretching region have been obtained from partially deuterated methyl, dimethyl and trimethyl germanes and arsines, ethyl germane, MeGeH2X, where X = F, Cl, Br, I and MeGeHCl2. In general, the isolated CH stretching frequencies show patterns of behaviour with respect to the α and β substituent effects of CH3, GeH3 and halogen groups similar to these found in the corresponding carbon compounds, although on a smaller scale. However, the trans effect of halogen is almost independent of the nature of the halogen, and a CH bond trans to GeH3 is weaker than one trans to CH3. In methyl arsine, an anomalous lone pair trans effect occurs, as previously found in methyl phosphine. CH bond lengths and dissociation energies, and average HCH angles are predicted for these compounds. 相似文献
10.
The crystal structure of tert-butoxycarbonyl-D, L-isoleucyl-glycyl methyl ester was determined by the X-ray method. The space group is P21 with a = 9.115(3), b= 12.075(4), c = 16.258(5) Å, β = 89.84(3) and Z = 4. It was observed that in the peptides which contain IleGly fragment the carbonyl group of Gly residue has less tendency to form hydrogen bond than that of Ile residue. This tendency was supported by the atomic net charges calculated with the quantum mechanics program CNDO/2. 相似文献
11.
Jzsef Emri 《Journal of Molecular Structure》2003,620(2-3):283-290
An exact relationship between bond length and bond order has been derived for the first time based on the concept of electron density. This relationship allows the calculation of sufficiently accurate bond orders and also determines the number of bond-forming electrons. According to this novel relationship between bond order and bond length, the bond order of the carbon–carbon bond in ethylene is 1.75, whereas it is 2.50 in acetylene. These bond orders are readily interpreted by the fragmentation of π-bonds and a consequent decrease in bond order, which is further supported by the chemical properties of these molecules. Assuming structure-specific fragmentation of π-bonds (i.e. one structural motif always adheres to one or two types of bond fragmentation scheme), the bond orders can be predicted for molecules containing multiple carbon–carbon bonds in excellent agreement with the experimental findings. 相似文献
12.
The versatility of palladium(II) acetate and palladium on activated charcoal catalysts with triethylsilane has been investigated in the hydrogenation and the isomerization of carbon–carbon double bond of 1‐alkenes. The reduction of 1‐alkenes was carried out in the presence of triethylsilane, ethanol and a catalytic amount of palladium(II) acetate or palladium on activated charcoal, at room temperature. This facile and efficient method affords high yields for hydrogenation of unsaturated alkenes to the corresponding alkanes. Then the carbon–carbon double bond isomerization of 1‐alkenes was tested using the same catalysts in the absence of solvent. The system palladium(II) acetate‐triethylsilane was found to be more effective compared with palladium on an activated charcoal–triethylsilane system at room temperature, while comparable results were obtained at 50 °C for both catalysts. Copyright © 2006 John Wiley & Sons, Ltd. 相似文献
13.
Qingzhong Li Hongjie Zhu Xiulin An Baoan Gong Jianbo Cheng 《International journal of quantum chemistry》2009,109(3):605-611
The nonadditivity of methyl group in the single‐electron hydrogen bond of the methyl radical‐water complex has been studied with quantum chemical calculations at the UMP2/6‐311++G(2df,2p) level. The bond lengths and interaction energies have been calculated in the four complexes: CH3? H2O, CH3CH2? H2O, (CH3)2CH? H2O, and (CH3)3C? H2O. With regard to the radicals, tert‐butyl radical forms the strongest hydrogen bond, followed by iso‐propyl radical and then ethyl radical; methyl radical forms the weakest hydrogen bond. These properties exhibit an indication of nonadditivity of the methyl group in the single‐electron hydrogen bond. The degree of nonadditivity of the methyl group is generally proportional to the number of methyl group in the radical. The shortening of the C···H distance and increase of the binding energy in the (CH3)2CH? H2O and (CH3)3C? H2O complexes are less two and three times as much as those in the CH3CH2? H2O complex, respectively. The result suggests that the nonadditivity among methyl groups is negative. Natural bond orbital (NBO) and atom in molecules (AIM) analyses also support such conclusions. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2009 相似文献
14.
Zhitao Zhang Na Zhao Kui Ma Qingpeng Cheng Jing Zhang Lirong Zheng Ye Tian Xingang Li 《中国化学快报》2019,30(2):513-516
Mordenite with the isolated zinc ions enhanced the catalytic performance for carbonylation of dimethyl ether to methyl acetate. The addition of Zn affected the acid properties of the catalysts, further changed the rate of coke deposition. 相似文献
15.
R. G. Sultangareev Yu. G. Kryazhev M. A. Kuznetsova 《Russian Chemical Bulletin》1969,18(10):2182-2183
Conclusions The reaction of ethynyl dimethyl carbinol and its methyl ether with trichlorobromomethane in the presence of benzoyl peroxide has given the corresponding adducts.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 10, pp. 2326–2327, October, 1969. 相似文献
16.
The variation of the fundamental hydroxyl stretching frequency in the infrared spectrum of a number of epoxy–diamine networks has been determined as a function of temperature. Short-range hydrogen bonds appear to be formed in all the epoxy networks examined. For polymers formed by using short-chain α,ω-diamines, it appears that a hydroxyl–hydroxyl bond is dominant. For longer chain lenghts, a hydroxylamine bond becomes important. The hydroxyl–hydroxyl bond breaks at the glass transition temperature, whereas the hydroxyl–amine bond breaks at a higher temperature. At the highest temperatures studied, all hydroxyl groups form weak short-range bonds to the alkyl-aryl ether function. The absorption frequency for the hydroxyl group in all types of hydrogen bonds formed shows a strong dependence on temperature and absorption moves to a higher wavenumber with increase in temperature. 相似文献
17.
Fourier transform ion cyclotron resonance mass spectrometry has been used to study the temperature and deuterium isotope effects on the methyl cation transfer reaction between protonated dimethyl ether and dimethyl ether to produce trimethyloxonium cation and methanol. From the temperature dependence of this bimolecular reaction it was possible to obtain thermodynamic information concerning the energy barrier for methyl cation transfer for the first time. From the slope of an Arrhenius plot, a value for DeltaH(++) of -1.1 +/- 1.2 kJ mol(-1) was obtained, while from the intercept a value for DeltaS(++) of -116 +/- 15 J K(-1) mol(-1) was derived. This yields a DeltaG(++)(298) value of 33.7 +/- 2.1 kJ mol(-1). All thermodynamic values were in good agreement with ab initio calculations. Rate constant ratios for the unimolecular dissociation forming trimethyloxonium cation and the dissociation re-forming reactants were extracted from the apparent bimolecular rate constant. Attempts at modeling the temperature dependence and isotope effects of the unimolecular dissociation forming trimethyloxonium cation were also made. 相似文献
18.
O. A. Raevskii N. G. Mumzhieva T. E. Kron E. N. Tsvetkov 《Russian Chemical Bulletin》1981,30(6):1026-1030
Conclusions The oxides of dimethyl(bromomethyl)- and dimethyl(iodomethyl)phosphines exist as equilibrium mixtures of conformer pairs in solution, the form predominating in nonpolar solvents being that in which the phosphoryl and carbon-halogen bonds are in trans orientation.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 6, pp. 1292–1296, June, 1981. 相似文献
19.
Medhanie Gebremedhin Gebru Radhey Shyam Yadav Hanan Teller Haya Kornweitz Palaniappan Subramanian Alex Schechter 《Journal of Energy Chemistry》2023,(8):454-464
In this work,the oxidation of a mixture of dimethyl ether(DME) and methyl formate(MF) was studied in both an aqueous electrochemical cell and a vapor-fed polymer electrolyte membrane fuel cell(PEMFC)utilizing a multi-metallic alloy catalyst,Pt3Pd3Sn2/C,discovered earlier by us.The current obtained during the bulk oxidation of a DME-saturated 1 M MF was higher than the summation of the currents provided by the two fuels separately,suggesting the cooperative effect... 相似文献