Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Time-Dependent Analytical Derivatives

Several attempts have been made to extend the analytical energy derivative method also to the case of time-dependent perturbations. The pseudo-energy derivative (BED) method of Rice and Handy (1991), the quasi-energy derivative (QED) method of Sasagane, et al., (1993) and the time-dependent second-order Mpller-Plesset method [Pg.248]

All these methods define response functions as derivatives of a perturbed time-dependent quasi-energy (Lowdin and Mukherjee, 1972 Langhoff et al, 1972 Kutzelnigg, 1992) [Pg.249]

At the SCF level all methods lead to the same expressions for the response functions as obtained in the random phase approximation, in Section 10.3, with the time-dependent Hartree-Fock approximation, in Chapter 11.1, or with SCF linear response theory. The QED and time-averaged QED method for an MCSCF energy was also shown to yield the same expressions as obtained from propagator or response theory in Sections 10.4 and 11.2. [Pg.250]

However, for non-variational wavefunctions and in particular at the MP2 level the various methods differ, despite the fact that they were constructed to give the correct static perturbation limit, meaning the same as obtained from taking derivatives of the time-independent MP2 energy in Section 12.2.2. The PED method started from the normal expression for the MP2 closed-shell energy, Eq. (9.68), but expressed with the time-dependent perturbed molecular orbitals from Eq. (11.32). In addition, it was required that the condition [Pg.250]

However, no time-dependent contribution was included in the first-order correlation coefficients apart from the time dependence of the molecular orbitals. [Pg.250]


We have given an analytical method of deriving a time-dependent solution to our problem that is complicated but illustrates an important method. Frequently, steady state solutions are all that is needed. [Pg.279]

For electron transfer processes with finite kinetics, the time dependence of the surface concentrations does not allow the application of the superposition principle, so it has not been possible to deduce explicit analytical solutions for multipulse techniques. In this case, numerical methods for the simulation of the response need to be used. In the case of SWV, a semi-analytical method based on the use of recursive formulae derived with the aid of the step-function method [26] for solving integral equations has been extensively used [6, 17, 27]. [Pg.485]

It has been proposed that there may be a close link between the amount of an element available to living matter and the fraction of the total content which is labile (with the lability value being loosely defined as the total, accessible, hydrated ion level). Either the whole or part of the analytical result may be derived from dissociation of labile complex ions or dissolution of moderately soluble compounds. If one or both of these two processes proceed at a relatively slow rate, the magnitude of the lability value becomes time dependent . Conversely, if a complex exchanges ligands fairly rapidly, the amount present in... [Pg.22]

These are obtained by introducing an explicit time dependence of the permittivity. This dependence, which is specific to each solvent is of a complex nature, cannot in general be represented through an analytic function. What we can do is to derive semiempirical formulae either by applying theoretical models based on measurements of relaxation times (such as that formulated by Debye) or by determining through experiments the behaviour of the permittivity with respect to the frequency of an external applied field. [Pg.122]

This chapter is organized as follows. In Section II, we show how quantum chaos systems can be controlled under the optimal fields obtained by OCT. The examples are a random matrix system and a quantum kicked rotor. (The former is considered as a strong-chaos-limit case, and the latter is considered as mixed regular-chaotic cases.) In Section III, a coarse-grained Rabi state is introduced to analyze the controlled dynamics in quantum chaos systems. We numerically obtain a smooth transition between time-dependent states, which justifies the use of such a picture. In Section IV, we derive an analytic expression for the optimal field under the assumption of the CG Rabi state, and we numerically show that the field can really steer an initial state to a target state in random matrix systems. Finally, we summarize the chapter and discuss further aspects of controlling quantum chaos. [Pg.437]

The retention times of analytes are controlled by the concentration(s) of the organic solvent(s) in the mobile phase. If a relatively small entropic contribution to the retention is neglected, theoretical considerations based either on the model of interaction indices [58], on the solubility parameter theory [51,52] or on the molecular statistical theory [57], lead to the derivation of a quadratic equation for the dependence of the logarithm of the retention factor of a solute. A, on the concentration of organic solvent. aqueous-organic mobile phase ... [Pg.40]

The first four terms of the function are commonly found in molecular mechanics strain energy functions, and they are modified Hooke s law functions. The last term has been added to insure the proper stereochemistry about asymmetric atoms. A model is refined by minimizing the highly nonlinear strain energy function with respect to the atomic coordinates. An adaptive pattern search routine is used for the strain energy minimization because it does not require analytical derivatives. The time necessary to obtain good molecular models depends on the number of atoms in the molecule, the flexibility of the structure, and the quality of the starting model. [Pg.112]

When the external electric field is time-dependent, there is no well-defined energy of the molecular system in accordance with Eq. (100), and the wave function response can thus not be retrieved from a variational condition on the energy as in the analytic derivative approach described above. Instead the response parameters have to be determined from the time-dependent Schrodinger equation, a procedure which was illustrated in Section 3 for the exact state case. In approximate state theories, however, our wave function space only partially spans the 7V-electron Hilbert space, and the response functions that correspond to an approximate state wave function will clearly be separate from those of the exact state wave function. This fact is disregarded in the sum-over-states approach, and, apart from the computational aspect of slowly converging SOS expressions, it is of little concern when highly accurate wave function models are used. But for less flexible wave function models, the correct response functions should be used in the calculation of nonlinear optical properties. [Pg.42]

Analytic response theory, which represents a particular formulation of time-dependent perturbation theory, has constituted a core technology in much of the this development. Response functions provide a universal representation of the response of a system to perturbations, and are applicable to all computational models, density-functional as well as wave-function models, and to all kinds of perturbations, dynamic as well as static, internal as well as external perturbations. The analytical character of the theory with properties evaluated from analytically derived expressions at finite frequencies, makes it applicable for a large range of experimental conditions. The theory is also model transferable in that, once the computational model has been defined, all properties are obtained on an equal footing, without further approximations. [Pg.152]

In principle, the differentiation is either done numerically in the so-called finite-field methods, or in an analytical scheme, or a combination of both. Numerical finite-field calculations are limited to derivatives with respect to static fields. Since SFG is an optical process that involves dynamic oscillating fields, it becomes necessary to use an analytical approach, such as the time-dependent Hartree Fock (TDHF) method. [Pg.373]

Inhibition of self-diffusion becomes apparent at timescales A short compared jD, where d is the characteristic pore diameter, because a fraction of molecules is always close to the walls, and their diffusion is hindered. This fraction depends on the surface-to-volume ratio (S/V) of the sample and the observation time A of the diffusion. This simple model leads to a fraction of 2DAf 5/V molecules being restricted in their motion. An analytical derivation [Mitl, Mit2, Mit6, Mit7] confirms an approximately linear dependence of the effective diffusion coefficient Deff(A) on... [Pg.188]

All redox titration curves we have discussed here are independent of the total analytical concentration C of the redox couple. (This is not always the case in the Cr2072 / Cr 3+couple the reduction of one Cr2072- generates two Cr3+ions, which leads to a concentration-dependent redox titration curve.) Therefore, the above expressions precisely give Cs/ln (10) times the first derivative of the progress curve of the corresponding redox titration. You can convince yourself that this is so in exercise 5.10-1. [Pg.218]


See other pages where Time-Dependent Analytical Derivatives is mentioned: [Pg.248]    [Pg.249]    [Pg.251]    [Pg.248]    [Pg.249]    [Pg.251]    [Pg.421]    [Pg.97]    [Pg.110]    [Pg.643]    [Pg.326]    [Pg.66]    [Pg.201]    [Pg.214]    [Pg.147]    [Pg.172]    [Pg.420]    [Pg.141]    [Pg.41]    [Pg.436]    [Pg.171]    [Pg.227]    [Pg.185]    [Pg.319]    [Pg.294]    [Pg.174]    [Pg.309]    [Pg.184]    [Pg.67]    [Pg.151]    [Pg.164]    [Pg.6]    [Pg.134]    [Pg.285]    [Pg.716]    [Pg.206]    [Pg.3002]    [Pg.105]    [Pg.45]    [Pg.19]    [Pg.705]   


SEARCH



Analytic derivatives

Analytical derivatives

© 2024 chempedia.info