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 polycarbosilanes (PCS) with meta-linkage bending unit ((SINGLE BOND)Me2Si(SINGLE BOND)m(SINGLE BOND)C6H4(SINGLE BOND)Me2Si(SINGLE BOND)CH2CH2(SINGLE BOND)) were successfully synthesized in mild conditions by hydrosilylation in the presence of [Pt{(CH2(DOUBLE BOND)CHSiMe2)2O}2]. The PCS obtained were soluble in various solvents owing to the lowering of the crystallinity. These properties are well compared with those of the PCS [(SINGLE BOND)Me2Si(SINGLE BOND)p(SINGLE BOND)C6H4(SINGLE BOND)Me2Si(SINGLE BOND)CH2CH2(SINGLE BOND)]n. © 1996 John Wiley & Sons, Inc.  相似文献   

2.
G2 ab initio calculations on all ABX three-membered rings (TMRs) that can be derived from cyclopropane by systematic substitution of the (SINGLE BOND)CH2 groups by (SINGLE BOND)NH or (SINGLE BOND)O groups have been performed. Our results show that the decrease in the A(SINGLE BOND)B bond length as the electronegativity of X increases is significantly larger than that found for the corresponding acyclic analogs. In general, a systematic substitution of the (SINGLE BOND)CH2 groups of cyclopropane by (SINGLE BOND)NH or (SINGLE BOND)O groups implies significant geometric changes that are not reflected in a parallel change of the corresponding conventional ring strain energy (CRSE). When the electronegativity of the groups forming the TMR increases the effect on the CRSE of the system is small, although the charge delocalization inside the ring decreases. The near constancy of the CRSE along the series can be explained in terms of the charge redistribution of the system where the (SINGLE BOND)CH2 groups play a crucial role. There are, however, significant changes in the hydrogenation energies of the TMR investigated; our results show that, when in an ABX three-membered ring, the electronegativity of X increases the hydrogenation energy of A(SINGLE BOND)B bond decreases and vice versa. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 1072–1086, 1998  相似文献   

3.
Extensive semiempirical SCF-MO calculations confirm that the exo-anomeric effect in methyl O-, N- and S-glycosides deals with an interaction of π-character along the C1(SINGLE BOND)Y1 bond in a X5(SINGLE BOND)C1(SINGLE BOND)Y1(SINGLE BOND)Me moiety (where X = O, S; Y = O, NH, S). The bond-order between orbitals of pπ symmetry on C1 and Y1 serves as a measure of all significant molecular orbital interactions responsible for the exo-anomeric stabilization. The set of simpler compounds X(SINGLE BOND)CH2(SINGLE BOND)Y (X = OH, SH, SeH, TeH; Y = OH, SH, SeH, TeH, NH2) on which the anomeric effect has been well studied was also calculated and it is noticeable that the π-bond-orders accord with the results of other analyses of the ab initio wave function accounting for the anomeric effect. Although the AM1 and the PM3 parameterizations of MNDO do not accurately reproduce the anomeric effect energetic, they do reproduce accordingly the expected variations in the molecular conformations of complex carbohydrates, and thus it follows that there are maximal π-bond-orders for the synclinal arrangement around the C1(SINGLE BOND)Y1 bond. In addition, the π-bond-orders show the same behavior for conformational preferences around the C1(SINGLE BOND)C′1 and the C5(SINGLE BOND)C6 bonds in methyl C-glycosides and in the hydroxymethyl group of α-D -glucose, respectively. © 1996 by John Wiley & Sons, Inc.  相似文献   

4.
The relative rate technique has been used to measure the hydroxyl radical (OH) reaction rate constant of ethyl 3-ethoxypropionate (EEP, CH3CH2(SINGLE BOND)O(SINGLE BOND)CH2CH2C(O)O(SINGLE BOND)CH2CH3). EEP reacts with OH with a bimolecular rate constant of (22.9±7.4)×10−12 cm3 molecule−1s−1 at 297±3 K and 1 atmosphere total pressure. In order to more clearly define EEP's atmospheric reaction mechanism, an investigation into the OH+EEP reaction products was also conducted. The OH+EEP reaction products and yields observed were: ethyl glyoxate (EG, 25±1% HC((DOUBLE BOND)O)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl (2-formyl) acetate (EFA, 4.86±0.2%, HC((DOUBLE BOND)O)(SINGLE BOND)CH2(SINGLE BOND)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl (3-formyloxy) propionate (EFP, 30±1%, HC((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH2(SINGLE BOND)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl formate (EF, 37±1%, HC((DOUBLE BOND)O)O(SINGLE BOND)CH2CH3), and acetaldehyde (4.9±0.2%, HC((DOUBLE BOND)O)CH3). Neither the EEP's OH rate constant nor the OH/EEP reaction products have been previously reported. The products' formation pathways are discussed in light of current understanding of oxygenated hydrocarbon atmospheric chemistry. © 1997 John Wiley & Sons, Inc.  相似文献   

5.
Minimum-energy structures of O2, CO, and NO iron–porphyrin (FeP) complexes, computed with the Car–Parrinello molecular dynamics, agree well with the available experimental data for synthetic heme models. The diatomic molecule induces a 0.3–0.4 Å displacement of the Fe atom out of the porphyrin nitrogen (Np) plane and a doming of the overall porphyrin ring. The energy of the iron–diatomic bond increases in the order Fe(SINGLE BOND)O2 (9 kcal/mol) < Fe(SINGLE BOND)CO (26 kcal/mol) < Fe(SINGLE BOND)NO (35 kcal/mol). The presence of an imidazole axial ligand increases the strength of the Fe(SINGLE BOND)O2 and Fe(SINGLE BOND)CO bonds (15 and 35 kcal/mol, respectively), with few structural changes with respect to the FeP(CO) and FeP(O2) complexes. In contrast, the imidazole ligand does not affect the energy of the Fe(SINGLE BOND)NO bond, but induces significant structural changes with respect to the FeP(NO) complex. Similar variations in the iron–imidazole bond with respect to the addition of CO, O2, and NO are also discussed. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 69: 31–35, 1998  相似文献   

6.
Novel oligomers possessing a backbone formed of ((TRIPLE BOND)Si(SINGLE BOND)CH2(SINGLE BOND)Si(TRIPLE BOND)) and (SINGLE BOND)Si(SINGLE BOND)n units were prepared by the copolycondensation of bis(chlorosilyl)methanes and various dichlorosilanes in the presence of sodium, in refluxing toluene. The effect of the respective molar ratios of comonomers on the yields and the structure of the copolymers was investigated. The role of substituents on silicon atoms in the ability of these materials to provide convenient ceramic precursors upon pyrolysis was examined. When (TRIPLE BOND)Si(SINGLE BOND)H bonds were present, thermal cross-linking was readily performed and ceramics possessing variable C/Si ratios were prepared.  相似文献   

7.
Three-membered ring (3MR) forming processes of X(SINGLE BOND)CH2(SINGLE BOND)CH2(SINGLE BOND)F and CH2(SINGLE BOND)C((SINGLE BOND)Y)(SINGLE BOND)CH2(SINGLE BOND)F (X(DOUBLE BOND)CH2, O, or S and Y(DOUBLE BOND)0 or S) through a gas phase neighboring group mechanism (SNi) are studied theoretically using the ab initio molecular orbital method with the 6–31+G* basis set. When electron correlation effects are considered, the activation (ΔG) and reaction energies (ΔG0) are lowered by ca. 10 kcal mol−1, indicating the importance of the electron correlation effect in these reactions. The contribution of entropy of activation (−TΔS) at 298 K to ΔG is very small, and the reactions are enthalpy controlled. The ΔG and ΔG0 values for these ring closure processes largely depend on the stabilities of the reactants and the heteroatom acting as a nucleophilic center. The Bell–Evans–Polanyi principle applies well to all these reaction series. © 1997 John Wiley & Sons, Inc. J Comput Chem 18 : 1773–1784, 1997  相似文献   

8.
The adsorptive properties of cyanide (CN) on coinage metal (M) electrodes (M=Cu, Ag, Au) have been investigated using a relativistic density functional method. The way to model the electrochemical potential applied to the electrodes is to consider the systems in the presence of a perturbative external field F. The field-perturbative approach is proven to be a suitable method in interpreting the observed spectral shifts with electrode potential. The calculated potential-dependent shifts of ωM(SINGLE BOND)CN and ωC(SINGLE BOND)M are similar for the three metals, in agreement with experiment observations. The relativistic effects are required to account for the similarity in the frequency shifts of ωM(SINGLE BOND)CN. The calculated vibrational tuning rates dωC(SINGLE BOND)N/dF are 6.61×10−7, 6.61×10−7, and 5.64×10−7 cm−1/(V/cm) for M=Cu, Ag, and Au, respectively. The coupling of the M(SINGLE BOND)CN and C(SINGLE BOND)N internal modes contributes significantly (about 25%) to the size of the frequency shifts ΔωC(SINGLE BOND)N of the ligand. The effect of electric fields on the metal(SINGLE BOND)CN bonding is also investigated. It is shown that changes in the magnitude of CN to the metal donation and M(SINGLE BOND)CN bond strength occur under the influence of the electric field. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 67: 175–185, 1998  相似文献   

9.
The interaction of an iron atom with molecular nitrogen was studied using density functional theory. Calculations were of the all-electron type and both conventional local and gradient-dependent models were used. A ground state of linear structure was found for Fe(SINGLE BOND)N2, with 2S + 1 = 3, whereas the triangular Fe(SINGLE BOND)N2 geometry, of C2v symmetry, was located 2.1 kcal/mol higher in energy, at least for the gradient-dependent model. The reversed order was found using the conventional local approximation. In Fe(SINGLE BOND)N2, the N(SINGLE BOND)N bond is strongly perturbed by the iron atom: It has a bond order of 2.4, a vibrational frequency of 1886 cm−1, and an equilibrium bond length of 1.16 Å: These values are 3.0, 2359 cm−1, and 1.095 Å, respectively, for the free N2 molecule. With the gradient-dependent model and corrections for nonsphericity of the Fe atom, a very small binding energy, 8.8 kcal/mol, was calculated for Fe(SINGLE BOND)N2. Quartet ground states were found for both Fe(SINGLE BOND)N+2 and Fe(SINGLE BOND)N2. The adiabatic ionization potential, electron affinity, and electronegativity were also computed; the predicted values are 7.2, 1.22, and 4.2 eV, respectively. © 1997 John Wiley & Sons, Inc.  相似文献   

10.
1,3,2-Benzimidazaborole, 1,3,2-benzoxaborole, and 1,3,2-benzothiazaborole were synthesized from the corresponding 2-benzazole N(SINGLE BOND)BH3 and 2-benzazole S(SINGLE BOND)BH3 adducts through a reductive transposition from the isolobal fragment X(SINGLE BOND)C(sp2) (DOUBLE BOND) N(sp2) (SINGLE BOND) B(sp3) (X (DOUBLE BOND) N, O, S) to the fragment X(SINGLE BOND)B(sp2) (DOUBLE BOND) N(sp2) (SINGLE BOND) C(sp3). N(SINGLE BOND)BH3 substitution shifts to lower frequencies 4-H, C-3a, and C-7a resonances. The X-ray diffraction analysis of 2-(o-methoxyphenyl)benzothiazole N(SINGLE BOND)BH3 adduct is reported. Two new tetracyclic boron-bridged compounds were observed as by-products (6,9-(ethyl)-diaza-2-oxa-1-bora[3,4,7,8]-dibenzobycyclo[4.3.0]-nona-3,7-diene, 6d, and 8-aza-9-oxa-2-thia-1-bora-[3,4,7,8]dibenzobycyclo[3.4.0]nona-3,7-diene, 15d, when 2-(o-methoxyphenyl)-1-ethylbenzimidazole-BH3 6b and 2-(o-methoxyphenyl)-benzothiazole-BH3 15b adducts were heated. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
Ammeline–melamine–formaldehyde resins (AMF Resins) containing 5–100 mol % of ammeline, were synthesized by polymerization of the preformed sodium salt of ammeline, melamine, and formaldehyde in basic medium by three methods. These resins, when cured, constitute a new class of thermosets. The rate of hydroxymethylation of the amino groups of the ammeline salt with formaldehyde was somewhat larger then that of the amino groups on melamine. At higher pH values ammeline insolubility was not a problem. The AMF resin composition was approximately equal to the mol ratio of the components originally charged. Both ammeline and melamine were consumed over the entire reaction period. Thus, it is possible to make approximately uniform random ammeline-melamine-formaldehyde resins (AMF) with any mol ratio of ammeline salt to melamine. By controlling the pH of the solution from which the resins were isolated, the (SINGLE BOND) O Na+/(SINGLE BOND) OH ratio could be varied. Resin melting points varied widely with the mol fraction of ammeline and the (SINGLE BOND) O Na+/(SINGLE BOND) OH ratio. AMF resin solubilities, shelf lives, cloud points, and water tolerance depended upon the method of preparation, pH, and other factors. The sodium salt of ammeline was hydroxymethylated in water more readily than ammeline. More highly methylolated ammeline species were readily formed in solution but upon precipitation only bis-N-hydroxymethylammeline was isolated. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
A comparative study of intermolecular potential energy curves is performed on the complexes H2O(SINGLE BOND)HF, H2O(SINGLE BOND)H2O, H2O(SINGLE BOND)H2S, and H2S(SINGLE BOND)H2S using nine different basis sets at the MP2 and DFT (BLYP and B3LYP) levels of theory. The basis set superposition error is corrected by means of the counterpoise scheme and based on the “chemical Hamiltonian approach.” The counterpoise and CHA-corrected DFT curves are generally close to each other. Using small basis sets, the B3LYP functional cannot be favored against the BLYP one because the BLYP results sometimes get closer to the MP2 values than those of B3LYP. From the results—including the available literature data—we suggest that one has to use at least polarized-valence triple-zeta-quality basis sets (TZV, 6-311G) for the investigation of hydrogen-bonded complexes. Special attention must be paid to the physical nature of the binding. If the dispersion forces become significant DFT methods are not able to describe the interaction. Proper correction for the basis set superposition error is found to be mandatory in all cases. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 575–584, 1998  相似文献   

13.
The ylidyl substituent of the chlorophosphane Ph3P(DOUBLE BOND)CAr(SINGLE BOND)PRCl exerts a strong influence on the P(SINGLE BOND)Cl bond. An X-ray structure investigation of the representative with Ar(DOUBLE BOND)Ph, R(DOUBLE BOND)Me reveals the longest P(SINGLE BOND)Cl bond ever observed for an acyclic chlorophosphine (226.2(1) pm). It is connected to a conformation that allows an effective negative hyperconjugation. The ylidyl chlorophosphanes with an amino group R are covalent in benzene but become more or less ionic in dichloromethane solution. The solvent-dependent dissociation equilibrium can be followed by 31P NMR spectra. In case of an enamine-derived ylidyl chlorophosphane, the equilibrium shifts almost completely from the covalent to the ionic side within a rather narrow range of solvent composition (20 to 70 vol % dichloromethane). © 1996 John Wiley & Sons, Inc.  相似文献   

14.
The dependence of the intramolecular charge transfer through the σ-electron systems of substituted hydrocarbons on the spatial arrangement of the X(SINGLE BOND)Cα and Cβ(SINGLE BOND)Cγ bonds has been studied using the perturbation theory for the one-electron density matrix (DM). Analytical expressions for the populations of the orbitals pertinent to the cis- and trans-bonds with respect to the X(SINGLE BOND)Cα bond have been obtained and analyzed. The Hamiltonian matrix elements determining the predominant direction of the above-defined charge transfer (cis or trans) have been revealed. The electron-accepting and electron-donating substituents (X) have been considered separately and the dependence of the resulting charge-transfer direction on the properties of substituent has been established. © 1996 John Wiley & Sons, Inc.  相似文献   

15.
Density functional calculations are reported for the molecular structures, harmonic vibrational frequencies, UV/visible spectra, and oxo-transferability of MoO2X2 (X = F, Cl, Br, I). Available experimental data have been used to check the validity of the theoretical calculations. Given the good agreement between theory and gas-phase experiment, predictions have been issued for the less studied members of this family of compounds. Furthermore, electronic spectra of the full series have been computed for the first time. For all transitions studied, excitation energies decrease in the order F > Cl > Br > I. Finally, the labilization of Mo(SINGLE BOND)O bonds generated by the HOMO(SINGLE BOND)LUMO transition, which is related to the oxygen-atom transfer reaction in the active site of molybdenum oxidoreductases, was also investigated. For MoO2Cl2 and MoO2Br2 compounds, the HOMO-LUMO transition yields a considerable lengthening of the Mo(SINGLE BOND)O bond, yet not requiring a large excitation energy. © 1997 John Wiley & Sons, Inc.  相似文献   

16.
A systematic quantum chemical study on the structure and stability of the major types of β-turn structures in peptides and proteins was performed at different levels of ab initio MO theory (MP2/6-31G*, HF/6-31G*, HF/3-21G) considering model turns of the general type Ac(SINGLE BOND)Xaa(SINGLE BOND)Yaa(SINGLE BOND)NHCH3 with the amino acids glycine, l - and d -alanine, aminoisobutyric acid, and l -proline. The influence of correlation effects, zero-point vibration energies, thermal energies, and entropies on the turn formation was examined. Solvent effects on the turn stabilities were estimated employing quantum chemical continuum approaches (Onsager's self-consistent reaction field and Tomasi's polarizable continuum models). The results provide insight into the geometry and stability relations between the various β-turn subtypes. They show some characteristic deviations from the widely accepted standard rotation angles of β turns. The stability order of the β-turn subtypes depends strongly on the amino acid type. Thus, the replacement of l -amino acids in the two conformation-determining turn positions by d - or α,α-disubstituted amino acid residues generally increases the turn formation tendency and can be used to favor distinct β-turn subtypes in peptide and protein design. The β-turn subtype preferences, depending on amino acid structure modifications, can be well illustrated by molecular dynamics simulations in the gas phase and in aqueous solution. © 1997 by John Wiley & Sons, Inc. J Comput Chem 18 : 1415–1430, 1997  相似文献   

17.
Density functional (DF) calculations of the tetrachloromethane cation and its most important competitive process, the formation of CCl+3, were carried out to explain the possible stability of CCl+4. From results obtained with B-LYP and B-P86 methods, it is possible to produce a slight Jahn-Teller (JT) effect for a Cs planar structure of the cation type CCl2(SINGLE BOND)Cl(SINGLE BOND)Cl+ compatible with the experimental data obtained by electron-spin resonance spectroscopy. A complex of C3v structure CCl+3(SINGLE BOND)Cl which is similar to the previous one found in CF+4 appears when symmetry-broken wave functions are used in HF-LYP and HF-P86 methods. Depending of the DF method employed, either one of the minima [Cs (planar) and C3v] is the most stable and competes with the dissociation of the molecular ion to give CCl+3. The JT stabilization energy is smaller when the JT active coordinates are considered. © 1997 John Wiley & Sons, Inc.  相似文献   

18.
In the present work, the relationship between the large substituent effects on 3J(C1H) in 1-X-3-M-bicyclo[1.1.1]pentanes, I , and the polarizability of the bridgehead C3(SINGLE BOND)Mα bond is investigated. The existence of such a relationship is suggested by the finding that the effect of an electronegative substituent X on 3J(C1Mα) couplings in I (M=H) is due to a distortion of the C3(SINGLE BOND)H bond toward the C1 center, which enhances the Fermi contact interaction. If such distortion originates in an electrostatic effect, then in other members of this series it can be expected that the substituent effects on 3J(C1Mα) couplings should depend strongly on the C3(SINGLE BOND)Mα bond polarizability. Two approaches are followed. First, the ab initio CLOPPA-IPPP method is applied to study the C3(SINGLE BOND)Mα bond contribution to the molecular static polarizability tensor in I (M=H, F, CH3). Such bond polarizabilities are found to follow the same trend as calculated as well as experimentally determined substituent effects on 3J(C1Mα) couplings, which were measured as part of this work in I [X=H, Cl; M=F, CH3 and X=OCH3; M=Sn(CH3)3]. Second, 3J(C1Mα) couplings (M=H, CH3) are calculated at an ab initio level for X=H, F, and they are compared with those obtained in the parent compound (X=H) if the calculation is carried out in the presence of an inhomogeneous electric field. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 181–188, 1998  相似文献   

19.
The simple and efficient a priori BSSE-free SCF method (CHA/F) proposed in the previous parts of this series is extended to the case of open-shell systems treated at the single-determinant UHF level of theory. The appropriate equations were derived and sample calculations are presented on three different (H3+(DOTTED BOND)H, H2O(DOTTED BOND)H, and CH4(DOTTED BOND)H) systems. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 68: 151–158, 1998  相似文献   

20.
Ab initio geometry optimizations were performed on gaseous protonated glycine using the second-order Møller–Plesset perturbation theory with the 6-31G*, 6-31G**, 6-31+G**, and 6-311+G** basis sets. Eight energy minima and 12 saddle points in the low-energy region of the electronic potential energy surface were characterized. The global minimum was an amino N-protonated conformer containing an ionic H bond between the (SINGLE BOND)NH3+ and O(DOUBLE BOND)C(DIAGONAL BOND)(DIAGONAL BOND) groups. The lowest energy O-protonated conformer was stabilized by a conjugative attraction between the nitrogen lone-pair electrons and the positively charged planar fragment (SINGLE BOND)C(OH)2+. Relative electronic energies of the nine N- and 11 O-protonated species fall in the ranges of 0–10 and 30–40 kcal mol−1. At room temperature the equilibrium distribution contained the most stable N-protonated conformer almost exclusively. Additional subjects for investigation include the effects of basis set and electron correlation on the predicted structures, nonbonded interactions that influence the relative stability of protonated conformers, conformational interconversions based on intrinsic reaction coordinate calculations, and kinetic pathways for protonation and associated changes in Gibbs free energy. The work provides geometric, energetic, and thermodynamic data pertinent to the study of gas-phase ion chemistry of amino acids and peptides. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 1862–1876, 1998  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