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 46 毫秒
1.
We study characteristic features of minimization of the Hartree-Fock-Roothaan energy with respect to nonlinear parameters of the Gaussian basis set. We describe and apply regularization of the discrete Newton-Raphson method based on the analysis of eigenvalues of the Hessian matrix. We discuss results of groundstate energy calculations for the molecules LiH, CH+, CH, He3 2+, BH2 +, and H2O in optimal ls-Gaussian basis sets. We find that, for molecules with four to six electrons, good accuracy is obtained with small basis sets consisting of ls-functions only.Translated from Teoreticheskaya i Éxperimental'naya Khimiya, Vol. 24, No. 2, pp. 215–218, March–April, 1988.  相似文献   

2.
Threshold photoelectron-photoion coincidence (TPEPICO) spectroscopy has been used to investigate the unimolecular chemistry of gas-phase methyl 2-methyl butanoate ions [CH3CH2CH(CH3)COOCH3·+]. This ester ion isomerizes to a lower energy distonic ion [CH2CH2CH(CH3)COHOCH3·+] prior to dissociating by the loss of C2H4. The asymmetric time of flight distributions, which arise from the slow rate of dissociation at low ion energies, provide information about the ion dissociation rates. By modeling these rates with assumed k(E) functions, the thermal energy distribution for room temperature sample, and the analyzer function for threshold electrons, it was possible to extract the dissociative photoionization threshold for methyl 2-methyl butanoate which at 0 K is 9.80 ± 0.01 eV as well as the dissociation barrier of the distonic ion of 0.86 ± 0.01 eV. By combining these with an estimated heat of formation of methyl 2-methyl butanoate, we derive a 0 K heat of formation of the distonic ion CH2CH2CH(CH3)COHOCH3·+ of 101.0 ± 2.0 kcal/mol. The product ion is the enol of methyl propionate, CH3CHCOHOCH3·+, which has a derived heat of formation at 0 K of 106.0 ± 2.0 kcal/mol.  相似文献   

3.
The MNDO method gives geometries for the molecular cations of organoberyllium compounds of types BeR2 and HBeR (R = CH3, CHCH2, CCH, CN, C5H5), of C4H4Be and CH3BeBeH3 and of the series CH4?n(BeH)n (n = 0–4) which have symmetries in precise accord with the predictions of the Jahn-Teller theorem. In the series CH4?n(BeH)n and CH4?n(BeH)n+, the barriers to inversion via a planar intermediate decrease with increasing n, are significantly smaller for the cations than for the neutral molecules, and are zero for CH(BeH)3+ and C(BeH)4+, both of which have their minimum energy when strictly planar at carbon.  相似文献   

4.
The main fragmentation pathways of the N-1, C-2 and C-4 stereoisomers of the 1,2-dimethyl-4-R-transdecahydroquinoline-4-ol N-oxides (R=C?CH, CH?CH2 and C2H5) under electron impact are discussed. The correlation between the mass spectrometric chromatographic behaviour and the configuration of polar groups in the N-oxides examined is discussed. The mass spectra of the N-1 stereoisomers may be subdivided into two groups, depending only on the orientation of N→O group and not of the 4-OH group. The spectra of N-oxides with the axial N-oxide group reveal less intense ions and much more intense [M? CH3]+, [M? O]+, [M? OH]+ and ions, whereas in the spectra of their equatorial epimers the abundance of the ions exceeds the intensities of the latter ions.  相似文献   

5.
Supersonically cooled jets of nitrogen, methane, ethane, cyclopropane, and azomethane are crossed with collimated streams of electrons. The CH (B2Σ? → X2Π) spectra resulting from the electron-induced dissociation of CH4, C2H6, and CH2)3 can be fit with rotation temperatures between 4000 and 6000 K for an electron energy of 100 eV. Flourescence spectra of N2+ (B2Σw+ → X2Π) from the dissociative ionization of azomethane yield a rotational temperature of =8×103 K; from ionization of molecular nitrogen the rotational temperature of B2Σw+ N2+ is 45 K. Mechanisms for these various processes are discussed.  相似文献   

6.
The [C4H6O] ion of structure [CH2?CHCH?CHOH] (a) is generated by loss of C4H8 from ionized 6,6-dimethyl-2-cyclohexen-1-ol. The heat of formation ΔHf of [CH2?CHCH?CHOH] was estimated to be 736 kJ mol?1. The isomeric ion [CH2?C(OH)CH?CH2] (b) was shown to have ΔHf, ? 761 kJ mol?1, 54 kJ mol?1 less than that of its keto analogue [CH3COCH?CH2]. Ion [CH2?C(OH)CH?CH2] may be generated by loss of C2H4 from ionized hex-1-en-3-one or by loss of C4H8 from ionized 4,4-dimethyl-2-cyclohexen-1-ol. The [C4H6O] ion generated by loss of C2H4 from ionized 2-cyclohexen-1-ol was shown to consist of a mixture of the above enol ions by comparing the metastable ion and collisional activation mass spectra of [CH2?CHCH?CHOH] and [CH2?C(OH)CH?CH2] ions with that of the above daughter ion. It is further concluded that prior to their major fragmentations by loss of CH3˙ and CO, [CH2?CHCH?CHOH]+˙ and [CH2?C(OH)CH?CH2] do not rearrange to their keto counterparts. The metastable ion and collisional activation characteristics of the isomeric allenic [C4H6O] ion [CH2?C?CHCH2OH] are also reported.  相似文献   

7.
Quadrupole mass spectrometry has been employed to characterize the ionic species in the discharges of pure CH4, CH4/H2, and CH4/Ar systems. For pure methane, the major positive ions in the discharge at low pressure (e.g., 0.15 torr) are CH 3 + , C2H 3 + , CH 2 + , C2H 2 + , CH 4 + , C2H 4 + , and C2H 5 + at high pressure (e.g., 0.5 torr) the major ions are CH 3 + , C2H 3 + , C2H 5 + , C3H 3 + , C H3H 7 + , C4H 7 + , C5H 7 + , C6H 5 + , and C7H 7 + . The relative abundances of C1 ions decrease with increasing pressure, whereas those of higher-order ions increase with pressure. For 5% CH4 + 95% H2 mixture, in addition to those sampling from the pure methane plasma at the lower pressure, H n + ions have also been detected. For 5% CH4 +95% Ar mixture, the principal ions are CH 3 + , CH 2 + , CH+, CH 5 + , Ar+, and ArH+; the ions containing more than two carbon atoms are negligible. In these discharges, the CH 3 + and C2H 3 + are the most important positive ions in C1 and C2 ions, respectively. The ions detected are believed to come from the sheath between the electrode and the luminous plasma, and have high kinetic energy. An ion-molecule reaction mechanism is proposed which can well explain the observed main features of ionic products.Died June 1, 1991.  相似文献   

8.
Raman spectra of the polycrystalline l-alanine analogs CH3CH(NH+3)COO?, CH3CH(ND+3)-COO?, CD3CD(NH+3)COO?, and CD3CD(ND+3)COO? have been obtained. A normal coordinate analysis is carried out based on the experimental frequencies of the four isotopic analogs and a 34 parameter valence-type force field defined in terms of local symmetry coordinates. The final refinement, in which five stretching force constants are constrained to fixed values obtained from bond length data, results in an average error of 7 cm?1 (0.9%) for the observed frequencies of the four isotopically substituted molecules. Band assignments are given in terms of the potential energy distribution for local symmetry coordinates. For non-deuterated l-alanine, the vibrations above 1420 cm?1 and below 950 cm?1 may be described as localized group vibrations. By contrast, the eight modes in the middle frequency range, viz. the three skeletal stretching, the COO? symmetric stretching, one NH+3 rocking, the symmetric CH3 deformation, and the two methyne CH deformation vibrations, are very strongly coupled to one another. Some decoupling appears to take place in the perdeutero molecule, and all but five modes can be described as localized group vibrations.  相似文献   

9.
We propose a new 13C DEPTQ+ NMR experiment, based on the improved DEPTQ experiment, which is designed to unequivocally identify all carbon multiplicities (Cq, CH, CH2, and CH3) in two experiments. Compared to this improved DEPTQ experiment, the DEPTQ+ is shorter and the different evolution delays are designed as spin echoes, which can be tuned to different 1JCH values; this is especially valuable when a large range of 1JCH coupling constants is to be expected. These modifications allow (i) a mutual leveling of the DEPT signal intensities, (ii) a reduction in J cross-talk in the Cq/CH spectrum, and (iii) more consistent and cleaner CH2/CH3 edited spectra. The new DEPTQ+ is expected to be attractive for fast 13C analysis of small-to medium sized molecules, especially in high-throughput laboratories. With concentrated samples and/or by exploiting the high sensitivity of cryogenically cooled 13C NMR probeheads, the efficacy of such investigations may be improved, as it is possible to unequivocally identify all carbon multiplicities, with only one scan, for each of the two independent DEPTQ+ experiments and without loss of quality.  相似文献   

10.
The ground state energies of CH 3 + , CH3, and CH 3 are calculated both in the SCF (near Hartree-Fock) approximation and in the IEPA-PNO scheme including correlation energy. Due to a more appropriate choice of the basis, our SCF-values for CH 3 are substantially better than previously published ones. Both CH 3 + and CH3 are planar whereas the equilibrium bond angles in CH 3 are nearly tetrahedral. The inversion barrier of CH 3 is 2kcal/mol. The force constants of the out-of-plane bending modes are changed by correlation in the case of CH3 from 0.03–1.8 mdyn/Å. The localized MO's that correspond to the CH-bonds are bent in the non-equilibrium geometries. The dependence of the different pair correlation contributions on the angle that describes out-of-plane deformation is analyzed. The electron affinity of CH3 is 0.3 eV. Finally the Pariser-Parr disproportionation reaction is analyzed in the light of the present results. Changes in correlation energy for this reaction amount to less than 1 eV.  相似文献   

11.
The preparations of CH2SF4 and CH3CHSF4 are presented and the structures are discussed. Addition reactions of polar species give a wide range of new compounds, like Hg(CH2SF5)2, F4AsCH2SF5, cisBrSF4CH3, cisF5SeOSF4CH2Br, a.o. While CH2SF4 decomposes at room temperature slowly to CH2CH2 and SF4, at high temperatures HF and CSF2 are formed. CH3CHSF4 gives mainly CH3CHF2 at room temperature. The “saturated” compounds CH3SF5 and C2H5SF5 have been prepared. They react with SbF5 in SO2 at low temperatures to form the cations CH3SF4+ and C2H5SF4+. The CH3SF4+ ion has been investigated in detail by nmr methods at low temperatures. It decomposes to CH3 and SF4, which react further in the SO2/SbF5 system to CH3OSO+ and SF3+.  相似文献   

12.
We present a systematic comparison of the correlation contribution at the level of the second-order polarization propagator approximation (SOPPA ) and MP 2 to the static dipole polarizability of (1) Be, BeH?, BH, CH+, MgH?, AIH, SiH+, and GeH+; (2) BH3, CH4, NH3, H2O, HF, BF, and F2; and (3) N2, CO, CN?, HCN, C2H2, and HCHO . Fairly extended basis sets were used in the calculations. We find that the agreement with experimental values is improved in SOPPA and MP .2 over the results at the SCF level. The signs and magnitudes of the correlation contribution in SOPPA are similar to those obtained in analytical derivative MP 2 calculations. However, it is not possible to say, in general, which method gives the largest correlation contribution or the best agreement with experiment, nor is it possible to make a priori prediction of the sign of the correlation contribution. For the first group of molecules, which have a quasi-degenerate ground state, additional CCDPPA and CCSDPPA calculations were performed and compared with polarizabilities obtained as analytical/numerical derivatives of the CCD and CCSD energies. The CCSDPPA results were found to be in better agreement with other calculations than were the SOPPA results, demonstrating the necessity of using methods based on infinite-order perturbation theory for these systems. © 1994 John Wiley & Sons, Inc.  相似文献   

13.
Solvent transports across the perfluorosulfonic acid-type membrane Flemion S were measured for aqueous electrolyte solutions under a temperature difference and under an osmotic pressure difference. H+, Li+, Na+, K+, NH 4 + , CH3NH 3 + , (CH3)2NH 2 + , (CH3)3NH+, (CH3)4N+, (C2H5)4N+, (n-C3H7)4N+ and (n-C4H9)4N+ were used as counterions. Water flux across the membrane in HCl solution is higher than that in the other electrolyte solutions because hydrogen ions can exchange with the hydrogen of the neighbor water molecules and contribute to the water transport across the membrane as a proton jump in conductivity. The direction of thermoosmosis across the membrane in HCl, NaCl, (CH3)4NCl and (C2H5)4NCl solutions was from the cold side to the hot side and that in LiCl, KCl, NH4Cl, CH3NH3Cl, (CH3)2NH2Cl and (n-C4H9)4NBr solutions was from the hot side to the cold side, although thermoosmosis across anion-exchange membranes always occurs toward the hot side.  相似文献   

14.
A series of esters RCOOR′ (where R, R′ = CH3, CH3CH2, (CH3)2CH, (CH3)3C) were reacted with the [(CH3CO)3]+ ion from biacetyl in an ion cyclotron resonance spectrometer. A steric effect influences the rate of formation of stable products [RCOOR′·CH3CO]+ and is used to determine that either oxygen of the ester may be initially acylated by [(CH3CO)3]+.  相似文献   

15.
Information on the solvation of thiolato complex cations [Co(en)2(SCH2COO)]+ [Co(en)2(SCH2CH(COO)NH2)]+, [Co(en)2(SCH2CH2NH2)]2+, sulfenato complexes [Co(en)2(SOCH2COO)]+ [Co(en)2{SOCH2CH(COO)NH2}]+, [Co(en)2(SOCH2CH2NH2)]2+, the sulfinato [Co(en)2{SO2CH2CH(COO)NH2}]+, [Co(en)2(SO2CH2CH2NH2)]2+ as well as of [Co(en)3]3+ has been obtained from solubility measurements in MeCN–H2O mixtures at 298.2 K. The single-ion Gibbs energies of transfer of the CoIII complexes were derived from the solubilities of picrate and perchlorate salts for the full range of MeCN–H2O mixtures. Single-ion Gibbs energies of transfer for the perchlorate ion are given. The effects of the solvent mixtures were interpreted in the framework of chemical bond formation between the ions and the individual solvent molecules.  相似文献   

16.
Summary We have applied a gauge origin invariant method for calculations of nuclear magnetic shielding constants to the singly bonded molecules BF, F2, BH3, CH4, NH3, H2O, and HF as well as to the1H shielding constants of HCN and C2H2. The calculations were performed at the RPA and second order polarization propagator (SOPPA) level of theory. For most molecules the correlation contribution in SOPPA is less diamagnetic than in the comparable MP2 calculations. For F2, SOPPA gives a large paramagnetic correlation correction whereas the MP2 method gives a very small correlation contribution. For all molecules agreement with experimental results is generally improved at the SOPPA level compared to RPA. We have also demonstrated that second order gauge origin invariant, common and local origin (SOLO) methods do not necessarily give the same shielding even in the limit of a converged basis set.  相似文献   

17.
The ionization potentials for the stereoisomers of trans-fused 1,2-dimethyl- and 1-ethyl-2-methyl-4-R-decahydroquinol-4-ols (R?C?CH, CH?CH2 or C2H5) and the appearance potentials for the [M–CH3]+ and [M–C2H5]+ ions (loss of 2-CH3 and 4-C2H5 groups potential, respectively) were measured by using the electron impact method. The ionization and appearance potential for [M–CH3]+ are always lower for the isomers with the axial 2-CH3 group. For the C-2 epimers, the difference between the appearance potentials for the [M–CH3]+ ion values is likely to be equal to the enthalpy differences between the ground states of the epimers and the dissociation energy differences between the axial and equatorial C2–CH3 bonds. The appearance potentials for [M–C2H5]+ for the C-4 epimers possessing the 4-C2H5 group were very similar. At the same time, the appearance potentials for the [M–CH3]+ ions were lower for less stable epimers which had an axial 4-C2H5 group.  相似文献   

18.
The relative energies of 11 [C3H3O]+ ions are calculated by different molecular orbital methods (MINDO/3, MNDO, ab initio with 3-21G and 4-31G* basis set and configuration interaction). The four most stable structures are: a ([CH2?CH? CO]+), b c ([CH?C? CHOH]+) and d ([CH2?C?COH]+); their relative energies at the CI/4-31G*//3-21G level are 0, 117, 171 and 218 kJ mol?1, respectively. The isomerizations c→[CH?CH? CHO]+→[CH2?C? CHO]+a and dissociations into [C2H3]++CO and [HCO]++C2H2 are explored. The calculated potential energy profile reveals that the energy-determining step is the 1,3-H migration c→[CH?CH? CHO]+. This explains the value of unity of the branching ratio and the spread of kinetic energy released for the two dissociation channels.  相似文献   

19.
The perturbative configuration interaction using strictly localized molecular orbitals, called the modified PCILO method, for which the use of the Rayleigh-Schrödinger many-body perturbation theory with the Moller-Plesset Hamiltonian partitioning is characteristic, has been proposed in this communication. On the CNDO/2 and INDO levels of Hamiltonian approximations strictly localized molecular orbitals have been constructed by solving modified Roothaan equations. From the zero and second order energy interatomic distances and harmonic force constants for some diatomic molecules have been calculated. The linear dependence of the correlation energy on the number of valence electrons in the series of the molecules CH4, CH3F, CH2F2, CHF3 and CF4 is perfect.  相似文献   

20.
The goals of the present study were (a) to create positively charged organo‐uranyl complexes with general formula [UO2(R)]+ (eg, R═CH3 and CH2CH3) by decarboxylation of [UO2(O2C─R)]+ precursors and (b) to identify the pathways by which the complexes, if formed, dissociate by collisional activation or otherwise react when exposed to gas‐phase H2O. Collision‐induced dissociation (CID) of both [UO2(O2C─CH3)]+ and [UO2(O2C─CH2CH3)]+ causes H+ transfer and elimination of a ketene to leave [UO2(OH)]+. However, CID of the alkoxides [UO2(OCH2CH3)]+ and [UO2(OCH2CH2CH3)]+ produced [UO2(CH3)]+ and [UO2(CH2CH3)]+, respectively. Isolation of [UO2(CH3)]+ and [UO2(CH2CH3)]+ for reaction with H2O caused formation of [UO2(H2O)]+ by elimination of ·CH3 and ·CH2CH3: Hydrolysis was not observed. CID of the acrylate and benzoate versions of the complexes, [UO2(O2C─CH═CH2)]+ and [UO2(O2C─C6H5)]+, caused decarboxylation to leave [UO2(CH═CH2)]+ and [UO2(C6H5)]+, respectively. These organometallic species do react with H2O to produce [UO2(OH)]+, and loss of the respective radicals to leave [UO2(H2O)]+ was not detected. Density functional theory calculations suggest that formation of [UO2(OH)]+, rather than the hydrated UVO2+, cation is energetically favored regardless of the precursor ion. However, for the [UO2(CH3)]+ and [UO2(CH2CH3)]+ precursors, the transition state energy for proton transfer to generate [UO2(OH)]+ and the associated neutral alkanes is higher than the path involving direct elimination of the organic neutral to form [UO2(H2O)]+. The situation is reversed for the [UO2(CH═CH2)]+ and [UO2(C6H5)]+ precursors: The transition state for proton transfer is lower than the energy required for creation of [UO2(H2O)]+ by elimination of CH═CH2 or C6H5 radical.  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