Big Chemical Encyclopedia

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

Articles Figures Tables About

Classical reaction probability

G.C. Schatz, J.M. Bowman, A. Kuppermann, Exact quantum, quasiclassical, and semi-classical reaction probabilities for the collinear F+TL -HF+F reaction, J. Chem. Phys. 63 (1975) 674. [Pg.158]

Lack of space prevents us from describing in any detail other properties of pods. It should be mentioned though that pods may be characterized as repulsive or attractive according to the local behaviour of trajectories in their vicinity. Loosely speaking, repulsive pods repel trajectories in their vicinity, attractive pods attract them. These properties have enabled the development of a lower bound to the classical reaction probability, the construction of a new theory of direct reactions as well as new numerical methods for computation of classical product state distributions.The repulsive attractive properties provide a global picture of the classical flow of trajectories at a given energy. [Pg.144]

At the time the experiments were perfomied (1984), this discrepancy between theory and experiment was attributed to quantum mechanical resonances drat led to enhanced reaction probability in the FlF(u = 3) chaimel for high impact parameter collisions. Flowever, since 1984, several new potential energy surfaces using a combination of ab initio calculations and empirical corrections were developed in which the bend potential near the barrier was found to be very flat or even non-collinear [49, M], in contrast to the Muckennan V surface. In 1988, Sato [ ] showed that classical trajectory calculations on a surface with a bent transition-state geometry produced angular distributions in which the FIF(u = 3) product was peaked at 0 = 0°, while the FIF(u = 2) product was predominantly scattered into the backward hemisphere (0 > 90°), thereby qualitatively reproducing the most important features in figure A3.7.5. [Pg.878]

The flux-flux expression and its extensions have been used to calculate reaction probabilities for several important reactions, including H2+02 H + H2O, by explicit calculation of the action of G in a grid representation with absorbmg potentials. The main power of the flux-flux fomuila over the long mn will be the natural way in which approximations and semi-classical expressions can be inserted into it to treat larger systems. [Pg.2303]

Aromatic denitrocyclizations have been used for many years in some well-known synthetic reactions. Probably the best known example is the Turpin synthesis of phenoxazines and similar synthesis of phenothiazines. The classical setup used usually base-catalyzed reactions in polar protic solvents, very often alcohols. In many cases using polar aprotic solvents was found advantageous. Besides the mentioned influence of the H-bonding, better ionization and lower solvation of the nucleophile are also important. Sf Ar reactions proceed through strongly polarized complexes, which are well soluble and highly polarized in polar aprotic solvents. [Pg.190]

Two commonly used synthetic methodologies for the synthesis of transition metal complexes with substituted cyclopentadienyl ligands are important. One is based on the functionalization at the ring periphery of Cp or Cp metal complexes and the other consists of the classical reaction of a suitable substituted cyclopentadienyl anion equivalent and a transition metal halide or carbonyl complex. However, a third strategy of creating a specifically substituted cyclopentadienyl ligand from smaller carbon units such as alkylidynes and alkynes within the coordination sphere is emerging and will probably find wider application [22]. [Pg.101]

In the adiabatic bend approximation (ABA) for the same reaction,18 the three radial coordinates are explicitly treated while an adiabatic approximation was used for the three angles. These reduced dimensional studies are dynamically approximate in nature, but nevertheless can provide important information characterizing polyatomic reactions, and they have been reviewed extensively by Clary,19 and Bowman and Schatz.20 However, quantitative determination of reaction probabilities, cross-sections and thermal reaction rates, and their relation to the internal states of the reactants would require explicit treatment of five or the full six degrees-of-freedom in these four-atom reactions, which TI methods could not handle. Other approximate quantum approaches such as the negative imaginary potential method16,21 and mixed classical and quantum time-dependent method have also been used.22... [Pg.412]

The key intermediate in the synthesis of the derivatives of 19-F3-androstane is the trifluoro analogue of the Wieland-Miescher ketone. Its preparation involves a Diels-Alder reaction between a trifluoromethyi ketone and a siloxy diene. Another original step is the regioselective reduction of a diketone only the ketone function in P of CF3 (probably activated by this substituent) is reduced (Figure 4.6). " Then, a succession of classical reactions leads to derivatives of androstane from the trifluoro analogue of the Wieland-Miescher ketone (Figure 4.7). ... [Pg.103]

In the classical version of the optical model, as developed mainly by Ross and co-workers,35-46-47 each impact parameter b manifests a certain reaction probability P(b), the classical opacity. Although a complex potential as in (11.12) is probably meaningless in a purely classical context, if we use the (rigorous) rate interpretation of T(r), we may derive24... [Pg.501]

The classical theory makes especially clear the inherent ambiguity of data analysis with the optical model, and this ambiguity carries over into the quantum model. If we wish to use experimental differential cross sections to gain information about V0(r) and P(b) or T(r), we must assume a reasonable parametric form for V0(r) that determines the shape of the cross section in the absence of reaction. The value P(b) is then determined [or T(r) chosen] by what is essentially an extrapolation of this parametric form. In the classical picture a V0(r) with a less steep repulsive wall yields a lower reaction probability from the same experimental cross-section data. The pair of functions V0 r), P b) or VQ(r), T(r) is thus underdetermined. The ambiguity may be relieved somewhat (to what extent is not yet known) by fitting several sets of data at different collision energies and, especially, by fitting other types of data such as total elastic and/or reactive cross sections simultaneously. [Pg.502]

As stated above, intermolecular coupling reactions between carbon atoms are of limited use. In the classical Wurtz reaction two identical primary alkyl iodide molecules are reduced by sodium. n-Hectane (C100H202), for example, has been made by this method in 60% yield (G. Stallberg, 1956). The unsymmetrical coupling of two alkyl halides can be achieved via dialkylcuprates. The first halide, which may have a branched carbon chain, is lithiated and allowed to react with copper(I) salts. The resulting dialkylcuprate can then be coupled with alkyl or aryl iodides or bromides. Although the reaction probably involves radicals it is quite stereoselective and leads to inversion of chiral halides. For example, lithium diphenyl-cuprate reacts with (R)-2-bromobutane with 90% stereoselectivity to form (S)-2-phenylbutane (G.M. Whitesides, 1969). [Pg.36]

Poliak, E., Child, M.S., and Pechukas, P, (1980). Classical transition state theory A lower bound to the reaction probability, J. Chem. Phys. 72, 1669-1678. [Pg.401]

The reaction cross-section or is related to the reaction probability, which can be calculated theoretically from the collision dynamics based on classical mechanics. [Pg.52]

Three-body (and many-body) quasi-classical scattering is formulated and the numerical evaluation of the reaction probability is described. [Pg.52]

A theoretical determination of the rate constant for a chemical reaction requires a calculation of the reaction cross-section based on the dynamics of the collision process between the reactant molecules. We shall develop a general relation, based on classical dynamics, between reaction probabilities that can be extracted from the dynamics of the collision process and the phenomenological reaction cross-section introduced in Chapter 2. That is, we give a recipe for how to calculate the reaction cross-section in accord with the general definition in Eq. (2.7). [Pg.53]

In this section, we shall consider how the solution of the classical equations of motion for more than two atoms may be used to find reaction probabilities and cross-sections for chemical reactions. Although the treatment is based on classical mechanics, it is termed quasi-classical because quantization of vibrational and rotational energy levels is accounted for. [Pg.74]

There is another more direct way of calculating the rate constant k(T), i.e., it is possible to bypass the calculation of the complete state-to-state reaction probabilities, S m(E) 2, or cross-sections prior to the evaluation of the rate constant. The formulation is based on the concept of reactive flux. We start with a version of this formulation based on classical dynamics and, in a subsequent section, we continue with the quantum mechanical version. It will become apparent in the next section that the classical version is valid not only in the gas phase, but in fact in any phase, that is, the foundation for condensed-phase applications will also be provided. [Pg.111]

Fig. 8.1.1 An illustration of the relations between the rate constants, k(T) and k(E), and the reaction probability P as obtained from either quantum mechanics, quasi-classical mechanics, or various assumptions (approximations) for the reaction dynamics. Fig. 8.1.1 An illustration of the relations between the rate constants, k(T) and k(E), and the reaction probability P as obtained from either quantum mechanics, quasi-classical mechanics, or various assumptions (approximations) for the reaction dynamics.
With complexes, the only dynamical calculations to date have been classical trajectories in which it was assumed that there is no significant C-Br chemical interaction following photoexcitation (Schatz and Fitzcharles 1988). Consequently, the role of the complex has been limited to H + C02 interactions sampled over the probability density for the intermolecular degrees of freedom, as well as the squeezed atom effect. These calculations have yielded reaction probabilities versus attack angle and nascent V, R, T excitations which are in reasonable agreement with the experimental results. Much work is still needed and the challenges are daunting. [Pg.84]

More recently, Yates and Lester230 fitted Liu s surface with a slightly modified form of the Porter-Karplus formulas after first fitting Liu s H2 potential to a simple Morse function. They then use the resulting surface to calculate the three-dimensional classical trajectory of the system. Their empirical fit very closely duplicates Liu s saddle-point properties. Reaction probabilities on this surface are compared with those on the PK surface. [Pg.54]

For any dynamical simulation, a continuous representation of the PES is mandatory since the potential and the gradients are needed for arbitrary configurations. One can in fact perform ab initio molecular dynamics simulations in which the forces necessary to integrate the classical equations of motion are determined in each step by an electronic structure calculations. There have been few examples for such an approach [35-37], However, in spite of the fact that electronic structure calculations can nowadays be performed very efficiently, still there is a significant numerical effort associated with ab initio calculations. This effort is so large that in the ab initio dynamics simulations addressing molecular adsorption and desorption at surfaces the number of calculated trajectories has been well below 100, a number that is much too low to extract any reliable reaction probabilities. [Pg.6]

In classical molecular dynamics simulations, reaction probabilities in general are determined by averaging over the results of many trajectories whose initial conditions are usually picked at random. The statistical uncertainty of the calculated reaction probabilities is then given by 1 /V N, where N is the number of calculated trajectories. This also means that it is computationally very demanding to determine small reaction probabilities since any calculated probability below 1 / JN is statistically not significant. [Pg.17]

In this context it is also worth mentioning that Showa Denko has developed a new process for the direct oxidation of ethene to acetic acid using a combination of palladium(II) and a heteropoly acid [104]. However, the reaction probably involves heteropoly acid-catalyzed hydration followed by palladium-catalyzed aerobic oxidation of ethanol to acetic acid rather than a classical Wacker mechanism. [Pg.161]


See other pages where Classical reaction probability is mentioned: [Pg.213]    [Pg.839]    [Pg.840]    [Pg.213]    [Pg.190]    [Pg.213]    [Pg.839]    [Pg.840]    [Pg.213]    [Pg.190]    [Pg.277]    [Pg.49]    [Pg.430]    [Pg.206]    [Pg.541]    [Pg.859]    [Pg.206]    [Pg.126]    [Pg.154]    [Pg.274]    [Pg.36]    [Pg.240]    [Pg.95]    [Pg.266]    [Pg.30]    [Pg.8]    [Pg.8]    [Pg.1514]   
See also in sourсe #XX -- [ Pg.170 ]




SEARCH



Semi-classical reaction probability

© 2024 chempedia.info