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 171 毫秒
1.
The features of the spectrum structure are considered for situations where some parameter λ of the N‐electron Hamiltonian reaches the threshold value η under which the discrete energy level falls into the continuous spectrum. The electron density properties are also studied. It is proved that for a sequence of the wave functions converging in energy to the lower bound of the continuous spectrum as λ approaches η the corresponding sequence of the electron densities converges to the density of the (N ? 1)‐electron ground state. The results generalize the Hellmann–Feynman theorem for the cases where only the one‐side energy derivatives exist or there is no limiting wave function. © 2002 Wiley Periodicals, Inc. Int J Quantum Chem, 2002  相似文献   

2.
The C3H6 has been investigated ab initio, taking all 24 electrons into account, using the Allgemeines Programmsystem/SCF ? MO ? LC (LCGO ) Methode. Variation of the C? C distance gives a total energy of ?116.02 a.u. at a C? C distance of 2.91 a.u. The ionization energy was found to be 10.33 eV.  相似文献   

3.
ACS symposium     
The SCF potential surface of the ground state for NO2 was calculated by using program JAMOL 3. The McLean–Loew–Berkowitz CGTO 's were used as basis functions. One of the two N? O distance R is fixed to 2.25 a.u. and the other one r and the ONO angle θ are varied from 2.25 to 5.0 a.u. and from 0° to 180°, respectively. The potential surface has the minimum around r = 2.50 a.u. and θ = 120°, where the energy is found to be ?203.954 a.u.  相似文献   

4.
The electronic structure of iron‐oxo porphyrin π‐cation radical complex Por·+FeIV?O (S? H) has been studied for doublet and quartet electronic states by means of two methods of the quantum chemical topology analysis: electron localization function (ELF) η(r) and electron density ρ(r). The formation of this complex leads to essential perturbation of the topological structure of the carbon–carbon bonds in porphyrin moiety. The double C?C bonds in the pyrrole anion subunits, represented by pair of bonding disynaptic basins Vi=1,2(C,C) in isolated porphyrin, are replaced by single attractor V(C,C)i=1–20 after complexation with the Fe cation. The iron–nitrogen bonds are covalent dative bonds, N→Fe, described by the disynaptic bonding basins V(Fe,N)i=1–4, where electron density is almost formed by the lone pairs of the N atoms. The nature of the iron–oxygen bond predicted by the ELF topological analysis, shows a main contribution of the electrostatic interaction, Feδ+···Oδ?, as long as no attractors between the C(Fe) and C(O) core basins were found, although there are common surfaces between the iron and oxygen basines and coupling between iron and oxygen lone pairs, that could be interpreted as a charge‐shift bond. The Fe? S bond, characterized by the disynaptic bonding basin V(Fe,S), is partially a dative bond with the lone pair donated from sulfur atom. The change of electronic state from the doublet (M = 2) to quartet (M = 4) leads to reorganization of spin polarization, which is observed only for the porphyrin skeleton (?0.43e to 0.50e) and S? H bond (?0.55e to 0.52e). © 2012 Wiley Periodicals, Inc.  相似文献   

5.
The time‐dependent quantum wave packet and the quasi‐classical trajectory (QCT) calculations for the title reactions are carried out using three recent‐developed accurate potential energy surfaces of the 11A′, 13A′, and 13A″ states. The two commonly used polarization‐dependent differential cross sections, dσ00/dωt, dσ20/dωt, with ωt being the polar coordinates of the product velocity ω′, and the three angular distributions, Pr), Pr), and Prr), with θr, Φr being the polar angles of the product angular momentum, are generated in the center‐of‐mass frame using the QCT method to gain insight into the alignment and the orientation of the product molecules. Influences of the potential energy surface, the collision energy, and the isotope mass on the stereodynamics are shown and discussed. Validity of the QCT calculation has been examined and proved in the comparison with the quantum wave packet calculation. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

6.
There is a contradiction as to the initial spatial separation ri of the two transient 2‐cyanoprop‐2‐yl radicals (Me2 ? CN) formed by flash photolysis of 2,2′‐azobis(isobutyronitrile) (AIBN) in solvents of various viscosities. The cage effect, expressed in terms of the in‐cage termination probability of the resulting radicals, is predicted correctly by classical Langevin models assuming a decrease of ri with increasing viscosity. However, the electron‐spin polarization of the radicals escaping the primary cage clearly indicates that the initial separation distance ri is independent of the solution viscosity. This obvious discrepancy can be reconciled by accounting for the strong electric dipole moments of these radicals and the resulting inter‐radical dipole? dipole interaction potential. We propose a primary‐caging model for polar radicals in solution based on an attractive inter‐radical mean‐force potential. The model is applied to the flash photolysis of AIBN and shown to describe properly the viscosity dependence of both the in‐cage termination probability (cage effect) and the electron‐spin polarization of the escaping 2‐cyanoprop‐2‐yl radicals.  相似文献   

7.
A set of spin-free functions ?i(r),i = 1 …? f, is obtained which form the basis of spin-free quantum chemistry. The ?i(r) show a one-to-one correspondence to antisymmetric space-spin functions Ψi(r, σ) with spin functions constructed according to Löwdin's projector operator method.  相似文献   

8.
We describe chemical bond changes as Franck–Condon electronic processes within a new theoretical ansatz that we call ‘rigged’ Born–Oppenheimer (R-BO) approach. The notion of the separability of nuclear and electron states implied in the standard Born–Oppenheimer (BO) scheme is retained. However, in the present scheme the electronic wave functions do not depend upon the nuclear coordinate (R-space). The new functions are obtained from an auxiliary Hamiltonian corresponding to the electronic system (r-coordinates) submitted to a Coulomb potential generated by external sources of charges in real space (α-coordinates) instead of massive nuclear objects. A stationary arrangement characterized by the coordinates α0A, is determined by a particular electronic wave function, ψ(r0A); it is only at this stationary point, where an electronic Schrödinger equation: He(r0A)|Ψ(r0A)=E0A)|Ψ(r0A) must hold. This equation permits us to use modern electronic methods based upon analytic first and second derivatives to construct model electronic wave functions and stationary geometry for external sources. If the set of wave functions {Ψ(r0A)} is made orthogonal, the energy functional in α-space, E(α;α0A)=Ψ(r0A)|He(r0A)|Ψ(r0A) is isomorphic to a potential energy function in R-space: E(R0A)=Ψ(r0A)|He(r,R)|Ψ(r0A). This functional defines, by hypothesis, a trapping convex potential in R-space and the nuclear quantum states are determined by a particular Schrödinger equation. The total wave function for the chemical species A reads as a product of our electronic wave function with the nuclear wave function (Ξik(R0A)): Φik(r,R)=Ψi(r0Aik(R0A). This approach facilitates the introduction of molecular frame without restrictions in the R-space. Two molecules (characterized with different electronic spectra) that are decomposable into the same number of particles (isomers) have the same Coulomb Hamiltonian and they are then characterized by different electronic wave functions for which no R-coordinate ‘deformation’ can possibly change its electronic structure. A bond breaking/forming process must be formally described as a spectroscopic-like electronic process. The theory provides an alternative to the adiabatic as well as the diabatic scheme for understanding molecular processes. As an illustration of the present ideas, the reaction of H2+CO leading to formaldehyde is examined in some detail.  相似文献   

9.
The AMO function of the hydrogen molecule ψ = ψc + η ψi, where ψc is the covalent part and ψi the ionic part, is investigated for small internuclear distances R. We found η → ?1 as R →,?1 as R → 0, contrary to the intuitively expected limit η → 1. However, near R = 0 an analytical expression of ψ is derived, showing that ψ reduces to the helium ground state as R → 0. We have proved that the empirical concept ?covalent and ionic character”? should be replaced by the symmetry argument in the case of small R.  相似文献   

10.
Two-dimensional fully numerical solutions of the Hartree–Fock problem are reported for the singlet ground states of H?, He, H2, and HeH+. The H2 energy at R = 1.4 a.u. is ?1.13362957 a.u.  相似文献   

11.
Floating spherical Gaussian orbital (FSGO ) open-shell calculations have been made to determine the potential energy surface of planar square and rectangular arrangements of the four-electron system H4. This surface is discussed in relation to the bimolecular isotope exchange reaction H2+D2-→ 2HD. The changes in energy and geometry accompanying the coplanar approach of two hydrogen molecules interacting chemically have also been investigated. Calculations on the electronic energies of planar T-shaped and kite arrangements of H4 of various sizes show that it is unlikely that these configurations can serve as transition states for the exchange reaction. However, the energy curve for linear configurations of H4 (H? H? H … H), calculated as a function of the H3 … H distance with the symmetric linear H3 (H-H-H) unit fixed at the internuclear distance of 1.9080 a.u., is found to have a deep minimum (?1.9176 a.u.) at an r(H3 … H) distance of 1.5846 a.u. The overall results suggest that the following mechanism for the exchange reaction, H2+H2→H2+H+H→H3+H→H+H2+ H→H2+H2 could be advantageous as it requires a barrier height of 0.1604 a.u. which is significantly lower than that calculated from the saddle point energy (0.1950 a.u.). However, the problem of reconciling this with the experimental activation energy of 0.0685 a.u. still remains.  相似文献   

12.
Ab initio multireference configuration interaction calculations for adiabatic potential curves, nonadiabatic couplings 〈φ i (R,r)|d/dR j (R,r)〉 and 〈φ i (R,r)|d2/dR 2 j (R,r)〉, and nuclear kinetic energy corrections 〈dφ i (R,r)/dR|dφ i (R,r)/dR〉 for the (3sσ) B and (3pσ) C1Σ+ Rydberg states of the CO molecule have been carried out. The energy positions and predissociation linewidths for the observed vibrational levels of these two states have been determined in a rigorous adiabatic representation by the complex scaling method employing a basis of complex scaled harmonic vibrational functions in conjunction with the Gauss-Hermite quadrature method to evaluate the complex Hamiltonian matrix elements. The present treatment correctly reproduces the observed trends in energies and line broadening for vibrational levels of the B1Σ+ state and represents an improvement over the previous treatment in literature. The errors in the determined spacings of the v = 0–4 vibrational levels of the C1Σ+ state are less than 2% compared with measured data. The predissociation linewidths for the v=3,4 levels of the C1Σ+ state are found to be 4.9 and 8.9 cm−1, respectively, in good agreement with the observed values. Received: 23 March 1998 / Accepted: 27 July 1998 / Published online: 9 October 1998  相似文献   

13.
Upper bounds are derived for |rμi|, μ = 1,2, ·, where ? denotes an exact electronic bound state wavefunction of a molecular system in the Born-Oppenheimer approximation, and ri is the distance of the ith electron from an appropriately chosen point, e.g., the molecular center. It is further shown that ? decays exponentially if ri → ∞.  相似文献   

14.
Formulas are derived for all Hamiltonian integrals required for molecular computations using a novel basis for single-center expansions. The basis orbitals depend exponentially upon α(r ? ρ)2 where r and ρ are the distance from center to electron and to a variationally scaled spherical shell, respectively. Comparisons are made between these so-called Gaussian shell orbitals (GSO ) and the conventional GTO and STO bases for single-center calculations. A preliminary comparison on H using a single GSO , a non-integer STO , and a GTO gives the optimized energies: ?0.51089 a.u., ?0.50504 a.u., and ?0.50422 a.u., respectively.  相似文献   

15.
Jacobsen  S.  Andresen  U.  Mäder  H. 《Structural chemistry》2003,14(2):217-225
The rotational spectra of o-fluorotoluene and its seven 13C isotopic species were recorded in the frequency range from 4 to 20 GHz with employment of pulsed molecular beam Fourier-transform microwave (MB-FTMW) spectrometers. The analysis of the spectra in the two lowest states of methyl internal rotation (torsional ground state, A and E species) was based on a asymmetric frame-rigid symmetric top Hamiltonian with inclusion of centrifugal distortion terms, yielding structural rotational constants, as well as the threefold barrier V 3 to internal rotation and the angle(a,i) between the principal moment of inertia a axis and the internal rotor axis i. The rotational constants of all eight isotopomeres were used to derive the seven 13C r s coordinates of the molecule.  相似文献   

16.
The thermal conductivity (λ) of carbon nanotubes (CNTs) with chirality indices (5,0), (10,0), (5,5), and (10,10) has been studied by reverse nonequilibrium molecular dynamics (RNEMD) simulations as a function of different bond length alternation patterns (Δri). The Δri dependence of the bond force constant (krx) in the molecular dynamics force field has been modeled with the help of an electronic band structure approach. These calculations show that the Δri dependence of krx in tubes with not too small a diameter can be mapped by a simple linear bond length–bond order correlation. A bond length alternation with an overall reduction in the length of the nanotube causes an enhancement of λ, whereas an alternation scheme leading to an elongation of the tube is coupled to a decrease of the thermal conductivity. This effect is more pronounced in carbon nanotubes with larger diameters. The formation of a polyene‐like structure in the direction of the longitudinal axis has a negligible influence on λ. A comparative analysis of the RNEMD and crystal orbital results indicates that Δri‐dependent modifications of λ and the electrical conductivity are uncorrelated. This behavior is in‐line with a heat transfer that is not carried by electrons. Modifications of λ as a function of the bond alternation in the (10,10) nanotube are explained with the help of power spectra, which provide access to the density of vibrational states. We have suggested longitudinal low‐energy modes in the spectra that might be responsible for the Δri dependence of λ. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

17.
For the one-electron Schrödinger equation among the solutions of which the Slater-Zener-type functions can be found, it is shown, that it can be generalized to the two-centre case only in one way, if one demands separability in prolate spheroidal coordinates, and if in addition to the Coulomb term of the potential energy there shall be an additional function of the product r 1 · r 2 only. The generalized problem with a potential energy of the form V(r) = ? Z1/r1 ? Z2/r2 ? Q(R)/r1r2 is studied for the case of two equal centres Z 1=Z2=Z≧0 with regard to the existence and number of bound states. The results are extended as far as possible also to the case with unequal centres. For some examples with equal centres wave functions and correlation diagrams have been computed exactly for the lowest electronic states.  相似文献   

18.
By using CASSCF/MRCI methods, theoretical molecular calculations have been performed for 12 electronic states for AlBr molecule and 12 electronic states for AlI molecule in the representation 2s+1Λ (neglecting spin‐orbit effects). Calculated potential energy curves are displayed. Spectroscopic constants including the harmonic vibrational wave number ωe, the electronic energy Te referred to the ground state and the equilibrium internuclear distance Re are predicted for these singlet and triplet electronic states for both AlBr and AlI molecules. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010  相似文献   

19.
The CH4 molecule has been investigated ab initio, for four different distances RC? H and one distortion of a triangle HCH about the equilibrium condition, taking all 10 electrons into account, using the Allgemeines Programmsystem/SCF ? MO ? LC (LCGO ) Methode. The equilibrium distance RC? H was estimated to 2.053 a.u., the minimum of the total energy to ?40.06 a.u., the heat of formation to 17.0 eV and the ionization energy to 14.8 eV. For the breathing force constant and the force constant of the twisting vibration a ω1 of 3139 cm?1 and a ω2 of 1865 cm?1 were found.  相似文献   

20.
《先进技术聚合物》2001,12(10):603-603
The original article to which this Erratum refers was published in Polym. Adv. Technol. 12 , 427–440 (2001) In the following paragraph, ‘µ’ was inserted instead of ‘γ’. The paragraph should have read: Xerographic Discharge The electric field (F) dependences of photogeneration efficiency as measured by the xerographic discharge method are shown in Figure 4 for PMPSi and PBMSi. The electric field and temperature (T) dependences of the photogeneration efficiency can be described using Onsager theory of geminate recombination (full lines). This theory can be used to describe one step in the photogeneration process, the thermal dissociation of ion‐pairs formed by light into free charge carriers in the external electric field. Under the assumption that the distribution of bound pairs is spherically symmetrical, the overall photogeneration efficiency ν is expressed as where ?0 is the voltage and temperature independent primary quantum yield (fraction of the ion‐pair generated per photon), g(r) is the spatial distribution of the pairs, and r is the pair separation distance. Function f(r,F,T) represents the Onsager dissociation probability [9] of the ion‐pair. The best theoretical fits of the experimental data given in Figure 4 were found using the Onsager theory with the Gaussian distribution of radii [17] ­ exp (?r22) using parameters: γ = 1.3 nm, η0 = 0.45 and γ = 1.6 nm, η0 = 0.85 for PMPSi and PBMSi, respectively. According to Eq. 2 the photogeneration quantum efficiency η is also temperature dependent. The activation energy was determined as Eη = 0.05 eV for PMPSi (inset in Figure 4) and 0.04 eV for PBMSi. The temperature dependence of the photogeneration efficiency is also in good agreement with the Onsager dissociation theory (full line in inset of Figure 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