Big Chemical Encyclopedia

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

Articles Figures Tables About

First-order reactions forces driving

The Houk group recently compared several artificial catalytic Diels-Alderase systems, including the ribozymes described here [11]. This study came to the conclusion that in none of these artificial systems there is a significant specific stabilization of the transition state. Acceleration arises predominantly from binding of the reactants, converting a second-order reaction of diene with dienophile into a first-order reaction of the termolecular complex of host, diene, and dienophile. The simultaneous presence of the two components of an intermolecular Diels-Alder reaction within the confined space of a cavity is the driving force that facilitates the reaction. [Pg.433]

For intermediate reaction rates the use of the enhancement factor is not consistent with the standard approach of diffusional limitations in reactor design and may be somewhat confusing. Furthermore, there are cases where there simply is no purely physical mass transfer process to refer to. For example, the chlorination of decane, which is dealt with in the coming Sec. 6.3.f on complex reactions or the oxidation of o-xylene in the liquid phase. Since those processes do not involve a diluent there is no corresponding mass transfer process to be referred to. This contrasts with gas-absorption processes like COj-absorption in aqueous alkaline solutions for which a comparison with C02-absorption in water is possible. The utilization factor approach for pseudo-first-order reactions leads to = tfikC i and, for these cases, refers to known concentrations C., and C . For very fast reactions, however, the utilization factor approach is less convenient, since the reaction rate coefficient frequently is not accurately known. The enhancement factor is based on the readily determined and in this case there is no problem with the driving force, since Cm = 0- Note also that both factors and Fji are closely related. Indeed, from Eqs. 6.3.C-5 and 6.3.C-10 for instantaneous reactions ... [Pg.320]

For reactions of order 1, the required reactor volume is greater for a backmixed reactor than for a plug flow one, e.g. by a factor of 191 for 90% conversion and 21.5 for 99% conversion, for a first order reaction. This is also true for gas-liquid mass transfer controlled systems, where in the main it is the gas phase residence time distribution which affects the driving force and hence reactor size for a given throughput. [Pg.353]

The dialytic regime is characterized by high surface reaction rate coefficients and by rate-limiting diffusion. The Sherwood number (Sh) characterizes the regimes. Sh is defined as the ratio of the driving force for diffusion in the boundary layer to the driving force for surface reaction alternatively, it is the ratio of the resistivity for diffusion to the resistivity for chemical reaction (reciprocal reaction rate coefficient). Diffusion limitation is the regime at Sh 1 and reaction limitation means Sh 1. The Sherwood number is closely related to the Biot, Nusselt, and Damkohler II numbers and the Thiele modulus. Some call it the CVD number. In the boundary-layer model it is a simple function of the thickness of the boundary layer, the diffusion coefficient, and the reaction rate coefficient. For simplicity a first-order reaction will be considered in the derivation below. [Pg.227]

The latter is a well-known quantity in the reaction-diffusion analysis in catalytic media (see Section 8.2.3) and can be written as the ratio between the average reaction rate over the washcoat cross-sectional area at a given axial position and its value at the surface. The former compares the driving force for mass transfer toward the coating, with the total potential for concentration decay (due to mass transfer and surface reaction). For a first-order reaction, 0 reduces to the Carberry number (see Chapter 3), and >/ is a concentration ratio between the averaged value inside the washcoat and the one at the surface. [Pg.200]

This difference is a measure of the free-energy driving force for the development reaction. If the development mechanism is treated as an electrode reaction such that the developing silver center functions as an electrode, then the electron-transfer step is first order in the concentration of D and first order in the surface area of the developing silver center (280) (Fig. 13). Phenomenologically, the rate of formation of metallic silver is given in equation 17,... [Pg.454]

Although the right-hand side of Eq. (14-60) remains valid even when chemical reactions are extremely slow, the mass-transfer driving force may become increasingly small, until finally c — Cj. For extremely slow first-order irreversible reactions, the following rate expression can be derived from Eq. (14-60) ... [Pg.1364]

Figure 11 Dependence on driving force of first-order rate constant for back electron transfer from colloidal Sn02 films to covalently attached complexes. The variations indicate that the reactions occur in the Marcus normal region. The identities of the molecular redox couples, listed from highest driving force to lowest, are, Rulll/n (5-Cl-phen)2 (phos-... Figure 11 Dependence on driving force of first-order rate constant for back electron transfer from colloidal Sn02 films to covalently attached complexes. The variations indicate that the reactions occur in the Marcus normal region. The identities of the molecular redox couples, listed from highest driving force to lowest, are, Rulll/n (5-Cl-phen)2 (phos-...
Figure 12 Plot of the first-order rate constant for back ET from high-area SnC>2 electrode to covalently bound RuIII(bpy)2(phosbpy)1 as a function of pH or H0. Notably, is nearly pH independent (ca. factor of 3 variations see Fig. 18). The solid line shows the behavior expected for an inverted reaction based on the 1.1-eV variation in driving force introduced by the pH variations. Figure 12 Plot of the first-order rate constant for back ET from high-area SnC>2 electrode to covalently bound RuIII(bpy)2(phosbpy)1 as a function of pH or H0. Notably, is nearly pH independent (ca. factor of 3 variations see Fig. 18). The solid line shows the behavior expected for an inverted reaction based on the 1.1-eV variation in driving force introduced by the pH variations.
The 1,3 "dipolar" redox isomerisation of C-captor-substituted amide chlorides is general, first order and independent of solvent polarity it is the fastest with the strongest captor groups (ref. 10). Since ionisation induced by counterion complexation of these amide chlorides blocks the isomerisation, either homolysis of CCl-bonds in opposition or a concerned elimination to the interceptable dipole and readdition of HC1 could explain the reaction mechanism. But what is the driving force of the rearrangement Again the Electronegativity Effect (ref. 2) is certainly... [Pg.206]

Step 1. Reactants enter a packed catalytic tubular reactor, and they must diffuse from the bulk fluid phase to the external surface of the solid catalyst. If external mass transfer limitations provide the dominant resistance in this sequence of diffusion, adsorption, and chemical reaction, then diffusion from the bulk fluid phase to the external surface of the catalyst is the slowest step in the overall process. Since rates of interphase mass transfer are expressed as a product of a mass transfer coefficient and a concentration driving force, the apparent rate at which reactants are converted to products follows a first-order process even though the true kinetics may not be described by a first-order rate expression. Hence, diffusion acts as an intruder and falsifies the true kinetics. The chemical kineticist seeks to minimize external and internal diffusional limitations in catalytic pellets and to extract kinetic information that is not camouflaged by rates of mass transfer. The reactor design engineer must identify the rate-limiting step that governs the reactant product conversion rate. [Pg.383]

In this example, if = 0.2 atm and = 0.004 atm, the driving force for mass transfer is R — R =0.196 atm. If the mass transfer coefficient is doubled—say, by increasing the velocity— the rate of mass transfer of A and the other rates would almost double. Assuming the surface reaction is first order in A, R would increase to about 0.008 atm, and the driving force for mass transfer would decrease slightly to 0.192 atm Thus, the rate would increase by a factor of 2 x 0.192/0.196 = 1.96. If, on the other hand, the kinetic constant for the surface reaction was doubled, the overall rate would... [Pg.50]


See other pages where First-order reactions forces driving is mentioned: [Pg.1094]    [Pg.96]    [Pg.442]    [Pg.412]    [Pg.226]    [Pg.450]    [Pg.258]    [Pg.228]    [Pg.114]    [Pg.389]    [Pg.292]    [Pg.337]    [Pg.114]    [Pg.39]    [Pg.1218]    [Pg.198]    [Pg.78]    [Pg.1979]    [Pg.904]    [Pg.114]    [Pg.400]    [Pg.83]    [Pg.108]    [Pg.652]    [Pg.90]    [Pg.199]    [Pg.1978]    [Pg.1218]    [Pg.4672]    [Pg.5262]    [Pg.660]    [Pg.228]    [Pg.210]    [Pg.35]   
See also in sourсe #XX -- [ Pg.87 ]




SEARCH



First reaction

First-order reactions

First-order reactions reaction

Reaction force

© 2024 chempedia.info