Big Chemical Encyclopedia

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

Articles Figures Tables About

Hamiltonian diatomic case

So far it has not proved possible, except in some very simple diatomic cases to calculate eigenfunctions of the Hamiltonian (16) using methods in which the nuclei... [Pg.108]

To generalize from the diatomic case, if the usual approach were taken to approximating solutions to the nuclear motion Hamiltonian using sums of products of electronic and nuclear parts a typical term in the siun used as trial function for the form (O Eq. 2.33) would be... [Pg.27]

The most commonly used semiempirical for describing PES s is the diatomics-in-molecules (DIM) method. This method uses a Hamiltonian with parameters for describing atomic and diatomic fragments within a molecule. The functional form, which is covered in detail by Tully, allows it to be parameterized from either ah initio calculations or spectroscopic results. The parameters must be fitted carefully in order for the method to give a reasonable description of the entire PES. Most cases where DIM yielded completely unreasonable results can be attributed to a poor fitting of parameters. Other semiempirical methods for describing the PES, which are discussed in the reviews below, are LEPS, hyperbolic map functions, the method of Agmon and Levine, and the mole-cules-in-molecules (MIM) method. [Pg.177]

Let us suppose that the system of interest does not possess a dipole moment as in the case of a homonuclear diatomic molecule. In this case, the leading term in the electric field-molecule interaction involves the polarizability, a, and the Hamiltonian is of the form ... [Pg.78]

In the general case R denotes a set of coordinates, and Ui(R) and Uf (R) are potential energy surfaces with a high dimension. However, the essential features can be understood from the simplest case, which is that of a diatomic molecule that loses one electron. Then Ui(R) is the potential energy curve for the ground state of the molecule, and Uf(R) that of the ion (see Fig. 19.2). If the ion is stable, which will be true for outer-sphere electron-transfer reactions, Uf(R) has a stable minimum, and its general shape will be similar to that of Ui(R). We can then apply the harmonic approximation to both states, so that the nuclear Hamiltonians Hi and Hf that correspond to Ui and Uf are sums of harmonic oscillator terms. To simplify the mathematics further, we make two additional assumptions ... [Pg.263]

To illustrate this point, consider a composite system composed of two noninteracting subsystems, one with p electrons (subsystem A) and the other with q = N — p electrons (subsystem B). This would be the case, for example, in the limit that a diatomic molecule A—B is stretched to infinite bond distance. Because subsystems A and B are noninteracting, there must exist disjoint sets Ba and Bb of orthonormal spin orbitals, one set associated with each subsystem, such that the composite system s Hamiltonian matrix can be written as a direct sum. [Pg.266]

There are many systems of different complexity ranging from diatomics to biomolecules (the sodium dimer, oxazine dye molecules, the reaction center of purple bacteria, the photoactive yellow protein, etc.) for which coherent oscillatory responses have been observed in the time and frequency gated (TFG) spontaneous emission (SE) spectra (see, e.g., [1] and references therein). In most cases, these oscillations are characterized by a single well-defined vibrational frequency, It is therefore logical to anticipate that a single optically active mode is responsible for these features, so that the description in terms of few-electronic-states-single-vibrational-mode system Hamiltonian may be appropriate. [Pg.303]

Relaxation times can be expressed in terms of time-correlation functions. Consider, for example, the case of a diatomic molecule relaxing from the vibrationally excited state n + 1> to the vibrational state /i> due to its interactions with a bath of solvent molecules. The Hamiltonian for the system is... [Pg.32]

Atoms in Molecules.—In this approach, which was first proposed by Moffitt,105 a wavefunction for a particular electronic state of a molecule is constructed from products of atomic wavefunctions, these, moreover, being taken to be exact eigenfunctions of their respective atomic hamiltonians. We confine our attention to the case of diatomic molecules AB so that, according to this procedure, the wavefunction is written as... [Pg.104]

It is well known from the Bom-Oppenheimer separation [1] that the pattern of energy levels for a typical diatomic molecule consists first of widely separated electronic states (A eiec 20000 cm-1). Each of these states then supports a set of more closely spaced vibrational levels (AEvib 1000 cm-1). Each of these vibrational levels in turn is spanned by closely spaced rotational levels ( A Emt 1 cm-1) and, in the case of open shell molecules, by fine and hyperfine states (A Efs 100 cm-1 and AEhts 0.01 cm-1). The objective is to construct an effective Hamiltonian which is capable of describing the detailed energy levels of the molecule in a single vibrational level of a particular electronic state. It is usual to derive this Hamiltonian in two stages because of the different nature of the electronic and nuclear coordinates. In the first step, which we describe in the present section, we derive a Hamiltonian which acts on all the vibrational states of a single electronic state. The operators thus remain explicitly dependent on the vibrational coordinate R (the intemuclear separation). In the second step, described in section 7.55, we remove the effects of terms in this intermediate Hamiltonian which couple different vibrational levels. The result is an effective Hamiltonian for each vibronic state. [Pg.316]

The Hamiltonian H(p,x,n, J ) is a model of diatomic molecules. h(p,x) represents the translational degrees of freedom and / ,(7t, Sj represents the internal vibrations of the molecules. If all the molecules are identical, we can assume that all frequencies are set to be equal. The internal part h n, E,) takes the form of uncoupled harmonic oscillators, so it looks specific. But this is not the case because all the nonlinear terms can be absorbed into the coupling term f P,X, 7I, ). [Pg.399]

The model Hamiltonian (13,6)—(13.8) and (13.13) and (13.14) can be used as a starting point within classical or quantum mechanics. For most diatomic molecules of interest hu> > ksT, which implies that our treatment must be quantum mechanical. In this case all dynamical variables in Eqs (13,6)—(13.8) and (13,13)—(13.14) become operators. [Pg.460]

This book was written to help spectroscopists understand the relationship between the exact molecular Hamiltonian, effective Hamiltonians used in fitting spectral data, and the molecular parameters obtained from both spectra and ab initio calculations. Although the general ideas for constructing effective Hamiltonians (Section 4.2) and several examples appropriate to special cases (for example the 2E+ 2n interaction in Sections 3.5.4 and 5.5) are discussed, no attempt is made here to present a complete and universal effective Hamiltonian for diatomic molecules. Brown, et al., (1979) derive an effective Hamiltonian that should be the starting point for the fitting of most non-1E, perturbation-free, diatomic molecular spectra. Other less general, effective Hamiltonians have been proposed, by De Santis, et al., (1973) for 3E states, by Brown and Milton (1976) for S > E-states, and by Brown and Merer (1979) for S > 1 n-states. [Pg.261]

The idea of an effective Hamiltonian for diatomic molecules was first articulated by Tinkham and Strandberg (1955) and later developed by Miller (1969) and Brown, et al., (1979). The crucial idea is that a spectrum-fitting model (for example Eq. 18 of Brown, et al., 1979) be defined in terms of the minimum number of linearly independent fit parameters. These fit parameters have no physical significance. However, if they are defined in terms of sums of matrix elements of the exact Hamiltonian (see Tables I and II of Brown, et al., 1979) or sums of parameters appropriate to a special limiting case (such as the unique perturber approximation, see Table III of Brown, et al., 1979, or pure precession, Section 5.5), then physically significant parameters suitable for comparison with the results of ab initio calculations are usually derivable from fit parameters. [Pg.261]

Radford (1961, 1962) and Radford and Broida (1962) presented a complete theory of the Zeeman effect for diatomic molecules that included perturbation effects. This led to a series of detailed investigations of the CN B2E+ (v — 0) A2II (v = 10) perturbation in which many of the techniques of modern high-resolution molecular spectroscopy and analysis were first demonstrated anticrossing spectroscopy (Radford and Broida, 1962, 1963), microwave optical double resonance (Evenson, et at, 1964), excited-state hyperfine structure with perturbations (Radford, 1964), effect of perturbations on radiative lifetimes and on inter-electronic-state collisional energy transfer (Radford and Broida, 1963). A similarly complete treatment of the effect of a magnetic field on the CO a,3E+ A1 perturbation complex is reported by Sykora and Vidal (1998). The AS = 0 selection rule for the Zeeman Hamiltonian leads to important differences between the CN B2E+ A2II and CO a/3E+ A1 perturbation plus Zeeman examples, primarily in the absence in the latter case of interference effects between the Zeeman and intramolecular perturbation terms. [Pg.418]

Consider the collision of an atom (denoted A) with a diatomic molecule (denoted BC), with motion of the atoms constrained to occur along a line. In this case there are two important degrees of freedom, the distance R between the atom and the centre of mass of the diatomic, and the diatomic intemuclear distance r. The Hamiltonian in terms of these coordinates is given by ... [Pg.970]


See other pages where Hamiltonian diatomic case is mentioned: [Pg.25]    [Pg.110]    [Pg.25]    [Pg.28]    [Pg.5]    [Pg.31]    [Pg.109]    [Pg.135]    [Pg.291]    [Pg.84]    [Pg.310]    [Pg.333]    [Pg.86]    [Pg.29]    [Pg.118]    [Pg.123]    [Pg.177]    [Pg.232]    [Pg.144]    [Pg.24]    [Pg.253]    [Pg.20]    [Pg.70]    [Pg.131]    [Pg.132]    [Pg.119]    [Pg.158]    [Pg.160]    [Pg.164]    [Pg.247]    [Pg.792]    [Pg.174]    [Pg.303]    [Pg.24]   
See also in sourсe #XX -- [ Pg.17 ]

See also in sourсe #XX -- [ Pg.17 ]




SEARCH



© 2024 chempedia.info