Big Chemical Encyclopedia

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

Articles Figures Tables About

Reaction constant approach, calculation

A complete description of any groundwater system necessitates consideration of reactions between rock forming minerals and the aqueous phase. This cannot be achieved without accurate thermodynamic properties of both the participating aluminosilicate minerals and aqueous aluminum species. Most computer codes used to calculate the distribution of species in the aqueous phase utilize the "reaction constant" approach as opposed to the "Gibbs free energy minimization" approach (3). In the former, aluminosilicate dissolution constants are usually written in terms of the aqueous aluminum species, Al, which is related to other aqueous aluminum species by appropriate dissociation reactions. [Pg.414]

It should be emphasized that the above equations, which relate reaction temperatures to calculated reactant or product energies, are equivalent to the more conventional linear free energy relationships, which relate logarithms of rate constants to calculated energies. It was felt that reactant temperatures would be more convenient to potential users of the present approach -those seeking possible new free radical initiators for polymerizations. [Pg.419]

Thus, given sufEcient detailed knowledge of the internal energy levels of the molecules participating in a reaction, we can calculate the relevant partition functions, and then the equilibrium constant from Eq. (67). This approach is applicable in general Determine the partition function, then estimate the chemical potentials of the reacting species, and the equilibrium constant can be determined. A few examples will illustrate this approach. [Pg.95]

In the quantum mechanical continuum model, the solute is embedded in a cavity while the solvent, treated as a continuous medium having the same dielectric constant as the bulk liquid, is incorporated in the solute Hamiltonian as a perturbation. In this reaction field approach, which has its origin in Onsager s work, the bulk medium is polarized by the solute molecules and subsequently back-polarizes the solute, etc. The continuum approach has been criticized for its neglect of the molecular structure of the solvent. Also, the higher-order moments of the charge distribution, which in general are not included in the calculations, may have important effects on the results. Another important limitation of the early implementations of this method was the lack of a realistic representation of the cavity form and size in relation to the shape of the solute. [Pg.334]

A second possible mathematical approach is to consider the equilibria via complexation or aggregation (21,22). Assuming that a drug compound can bind several tenside molecules in steps (concentration gradient), the brutto equilibria constants or stability constants KA can be given as follows. Normally the aggregation constant is calculated for 1 1 reactions or other known stoichiometric ratios (z) ... [Pg.122]

As the styrene process shows, it is not generally feasible to operate a reactor with a conversion per pass equal to the equilibrium conversion. The rate of a chemical reaction decreases as equilibrium is approached, so that the equilibrium conversion can only be attained if either the reactor is very large or the reaction unusually fast. The size of reactor required to give any particular conversion, which of course cannot exceed the maximum conversion predicted from the equilibrium constant, is calculated from the kinetics of the reaction. For this purpose we need quantitative data on the rate of reaction, and the rate equations which describe the kinetics are considered in the following section. [Pg.15]

Polarography is valuable not only for studies of reactions which take place in the bulk of the solution, but also for the determination of both equilibrium and rate constants of fast reactions that occur in the vicinity of the electrode. Nevertheless, the study of kinetics is practically restricted to the study of reversible reactions, whereas in bulk reactions irreversible processes can also be followed. The study of fast reactions is in principle a perturbation method the system is displaced from equilibrium by electrolysis and the re-establishment of equilibrium is followed. Methodologically, the approach is also different for rapidly established equilibria the shift of the half-wave potential is followed to obtain approximate information on the value of the equilibrium constant. The rate constants of reactions in the vicinity of the electrode surface can be determined for such reactions in which the re-establishment of the equilibria is fast and comparable with the drop-time (3 s) but not for extremely fast reactions. For the calculation, it is important to measure the value of the limiting current ( ) under conditions when the reestablishment of the equilibrium is not extremely fast, and to measure the diffusion current (id) under conditions when the chemical reaction is extremely fast finally, it is important to have access to a value of the equilibrium constant measured by an independent method. [Pg.26]

In Chapter 5, attention is directed toward the direct calculation of k(T), i.e., a method that bypasses the detailed state-to-state reaction cross-sections. In this approach the rate constant is calculated from the reactive flux of population across a dividing surface on the potential energy surface, an approach that also prepares for subsequent applications to condensed-phase reaction dynamics. In Chapter 6, we continue with the direct calculation of k(T) and the whole chapter is devoted to the approximate but very important approach of transition-state theory. The underlying assumptions of this theory imply that rate constants can be obtained from a stationary equilibrium flux without any explicit consideration of the reaction dynamics. [Pg.385]

This approach was successfully used in modeling the CVD of silicon nitride (Si3N4) films [18, 19, 22, 23]. Alternatively, molecular dynamics (MD) simulations can be used instead of or in combination with the MC approach to simulate kinetic steps of film evolution during the growth process (see, for example, a study of Zr02 deposition on the Si(100) surface [24]). Finally, the results of these simulations (overall reaction constants and film characteristics) can be used in the subsequent reactor modeling and the detailed calculations of film structure and properties, including defects and impurities. [Pg.469]

Different research groups have different approaches to elucidating decomposition pathways. Thus, a wide variety of activation parameters have been published (Table 1). This is a result of examining the decomposition on different timescales or temperatures. Often if the rate constants are calculated for a common temperature, they will be similar in magnitude [97], It is rare that the first paper published about a compound tells the entire story. For example, the papers on nitramine decomposition, cited herein, span thirty-five years. Each reveals a different piece of the story or supports previous postulates. Consider the mechanisms illustrated by Figs. 8-10. These are the results of several researchers approaching the problem from different perspectives yet, they make similar conclusions. As new ways of probing reaction chemistry become available, they should be implemented. At the same time the reaction chemistry should be examined by the more conventional techniques, and an effort should be made to correlate the results. [Pg.34]

Rate constants for diffusion-controlled reactions can be calculated from the laws of diffusion [18, 869]. For a simple cage reaction A -I- B AB, in which A reacts with B every time the two approach one another to within a distance R, the following equation can be derived,... [Pg.304]

Various solvent effect theories concerning HFS constants in ESR spectra using various reaction held approaches have been developed by Reddoch et al. [385] and Abe et al. [392]. According to Reddoch et al., none of the continuum reaction held models is entirely satisfactory. Therefore, a dipole-dipole model using a held due only to a dipole moment of one solvent molecule instead of various reaction fields was proposed, and applied to di-t-butyl aminyloxide [385]. However, Abe et al. found that the HFS constants are proportional to the reaction held of Block and Walker [393] when protic solvents are excluded [392]. This relationship has been successfully applied to di-t-butyl and diaryl aminyloxides, to the 4-(methoxycarbonyl)-l-methylpyridinyl radical cf. Fig. 6-10), and to the 4-acetyl-1-methylpyridinyl radical (see below) [392]. For another theoretical approach to the calculation of gr-values and HFS constants for di-t-butyl aminyloxide, see reference [501]. [Pg.373]

Table 1. Equations for calculating rate constants for simple reaction mechanisms based on the reaction order approach for DPSC [137],... Table 1. Equations for calculating rate constants for simple reaction mechanisms based on the reaction order approach for DPSC [137],...
The pairwise Brownian dynamics method has several advantages over numerical methods based on Smoluchowski s [9] approach (e.g., finite element method), and we discuss these here. The primary advantage of the method is the ease of mathematical formulation even for cases involving complex reaction site geometries, hydrodynamic interactions, charge effects, anisotropic diffusion and flow fields. Furthermore the method obviates the need to solve complex diffusion equations to obtain the concentration field from which the rate constant is calculated in the Smoluchowski method. In contrast, the rate constant is obtained directly in the pairwise Brownian dynamics method. The effective rate constants for different reaction conditions may be obtained from a single simulation this is not possible using the finite element method. [Pg.821]

Chemical equilibrium dictates the extent of reaction possible for a given temperature and pressure. For single simple reactions, an equilibrium constant approach can be used to determine the equilibrium concentration of gases for a given reaction. At equilibrium, the forward and the reverse reaction rates are equal. The equilibrium constant is calculated from the Gibbs free energy, as follows ... [Pg.169]

Hwang et al.131 were the first to calculate the contribution of tunneling and other nuclear quantum effects to enzyme catalysis. Since then, and in particular in the past few years, there has been a significant increase in simulations of QM-nuclear effects in enzyme reactions. The approaches used range from the quantized classical path (QCP) (e.g., Refs. 4,57,136), the centroid path integral approach,137,138 and vibrational TS theory,139 to the molecular dynamics with quantum transition (MDQT) surface hopping method.140 Most studies did not yet examine the reference water reaction, and thus could only evaluate the QM contribution to the enzyme rate constant, rather than the corresponding catalytic effect. However, studies that explored the actual catalytic contributions (e.g., Refs. 4,57,136) concluded that the QM contributions are similar for the reaction in the enzyme and in solution, and thus, do not contribute to catalysis. [Pg.298]

This approach was applied to the determination of silica surface distribution functions on reaction constants of organosilicon compounds chemisorption using the Gamma distribution on k (Table 4). The distribution functions on Ink were calculated on the basis of chemisorption kinetic isotherms for methylchlorosilanes (CH3)4 nSiCln (n from 1 to 4) and for the compounds (CH3)3SiX X = Cl, CN, N3, NCO, NCS, 0Si(CH3)3 on the non-porous dehydrated silica surface [4,22]. As shown in [57] the correlation is observed between the logarithms of rate constants with the sum of inductive constants of substituents at Si atom in (CH3)4 nSiCln compounds ([Pg.269]

Various chemical surface complexation models have been developed to describe potentiometric titration and metal adsorption data at the oxide—mineral solution interface. Surface complexation models provide molecular descriptions of metal adsorption using an equilibrium approach that defines surface species, chemical reactions, mass balances, and charge balances. Thermodynamic properties such as solid-phase activity coefficients and equilibrium constants are calculated mathematically. The major advancement of the chemical surface complexation models is consideration of charge on both the adsorbate metal ion and the adsorbent surface. In addition, these models can provide insight into the stoichiometry and reactivity of adsorbed species. Application of these models to reference oxide minerals has been extensive, but their use in describing ion adsorption by clay minerals, organic materials, and soils has been more limited. [Pg.220]

Equations (10-6) and (10-7) show that for the intermediate case the observed rate is a function of both the rate-of-reaction constant, ic and.. the mass-transfer coefficient k. In a design problem k and k would be known, so that Eqs. (10-6) and (10-7) give the global rate in terms of Cj. Alternately, in interpreting laboratory kinetic data k would be measured. If k is known, k can be calculated from Eq. (10-7). In the event that the reaction is not first order Eqs. (10-1) and (10-2) cannot be combined easily to eliminate C. The preferred approach is to utilize the mass-transfer coefficient to evaluate Q and then apply Eq. (10-2) to determine the order of the reaction n and the numerical value of k. One example of this approach is described by Olson et al. ... [Pg.361]

EROS handles concurrent reactions with a kinetic modeling approach, where the fastest reaction has the highest probability to occur in a mixture. The data for the kinetic model are derived from relative or sometimes absolute reaction rate constants. Rates of different reaction paths are obtained by evaluation mechanisms included in the rule base that lead to partial differential equations for the reaction rate. Three methods are available that cover the integration of the differential equations the GEAR algorithm, the Runge-Kutta method, and the Runge-Kutta-Merson method [120,121], The estimation of a reaction rate is not always possible. In this case, probabilities for the different reaction pathways are calculated based on probabilities for individual reaction steps. [Pg.233]

However, some further discussion is in order. In spite of Figure 6, it should not yet be generalized that the water gas shift reaction, under the influence of an ozonizer discharge, approaches thermodynamic equilibrium. Lunt has indicated that the equilibrium constant numbers calculated from his steady-state compositions vary with pressure. Our runs 176, 161, and 186 in Table III, while not at steady-state, indicate that this is possible. Further conversion vs. time experiments at various pressures are required. [Pg.233]

The rate constants were calculated with the transition state theory (TST) for direct abstraction reactions and the Rice-Ramsperger-Kassel-Marcus (RRKM) theory for reactions occuring via long-lived intermediates. For reactions taking place without well-defined TS s, the Variflex [35] code and the ChemRate [36] code were used for one-well and multi-well systems, respectively, based on the variational transition-state theory approach... [Pg.377]

In principle, the free energy change and in turn the equilibrium constant for such reactions can be calculated from conventional thermodynamic data (molar enthalpy, entropy, volume data) on the end-member isotopic species denoted in the reaction. This approach, however, is generally not practicable because of the paucity of thermodynamic data on isotopically pure end-members. Moreover, even if such data were widely available, the Gibbs free energy changes associated with most isotope exchange reactions... [Pg.2]

The slope of the Gibbs energy with respect to represents the Gibbs energy of reaction. It is zero as the reaction system approaches the equilibrium composition at constant p and T. This relation is used as the calculation basis for determining equilibria, especially for complex systems consisting of a large number of components and phases. [Pg.1951]

Ans. In the cell, glucose is phosphorylated in an enzymatic reaction with ATP. To see how the cell accomplishes this apparently unfavorable reaction, we must calculate the equilibrium constant for the cellular reaction. The equilibrium constant for the hydrolysis of ATP is well known, so that our best approach is to consider the reaction of ATP with glucose as consisting of two reactions with a common intermediate, listing their equilibrium constants along with the reactions ... [Pg.456]

Representative values of (3) and (4) are tabulated in Table I. Although the individual concentrations vary widely, each of these ratios is constant within experimental error at different heights above the burner, indicating that each of the reactions is equilibrated. Calculation of Keq is more uncertain for Reaction 2 because of the uncertain heat of formation of NH. Given this uncertainty, one can vary AHfl[NH) until fi obtained from Reaction 2 agrees with that obtained from Reaction 1. Such an approach yields fi = 5.05 x 10 with AHf (NH) = 89 kcal/mole. [Pg.75]


See other pages where Reaction constant approach, calculation is mentioned: [Pg.374]    [Pg.194]    [Pg.3]    [Pg.10]    [Pg.463]    [Pg.205]    [Pg.34]    [Pg.87]    [Pg.87]    [Pg.128]    [Pg.581]    [Pg.262]    [Pg.17]    [Pg.186]    [Pg.54]    [Pg.451]    [Pg.17]    [Pg.151]    [Pg.178]    [Pg.69]    [Pg.30]    [Pg.536]   


SEARCH



Calculations reactions

Constants calculated

© 2024 chempedia.info