Big Chemical Encyclopedia

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

Articles Figures Tables About

Diffusion and bimolecular reactions

Photosensitization of diaryliodonium salts by anthracene occurs by a photoredox reaction in which an electron is transferred from an excited singlet or triplet state of the anthracene to the diaryliodonium initiator.13"15,17 The lifetimes of the anthracene singlet and triplet states are on the order of nanoseconds and microseconds respectively, and the bimolecular electron transfer reactions between the anthracene and the initiator are limited by the rate of diffusion of reactants, which in turn depends upon the system viscosity. In this contribution, we have studied the effects of viscosity on the rate of the photosensitization reaction of diaryliodonium salts by anthracene. Using steady-state fluorescence spectroscopy, we have characterized the photosensitization rate in propanol/glycerol solutions of varying viscosities. The results were analyzed using numerical solutions of the photophysical kinetic equations in conjunction with the mathematical relationships provided by the Smoluchowski16 theory for the rate constants of the diffusion-controlled bimolecular reactions. [Pg.96]

The bulk polymerization of acrylonitrile in this range of temperatures exhibits kinetic features very similar to those observed with acrylic acid (cf. Table I). The very low over-all activation energies (11.3 and 12.5 Kj.mole-l) found in both systems suggest a high temperature coefficient for the termination step such as would be expected for a diffusion controlled bimolecular reaction involving two polymeric radicals. It follows that for these systems, in which radicals disappear rapidly and where the post-polymerization is strongly reduced, the concepts of nonsteady-state and of occluded polymer chains can hardly explain the observed auto-acceleration. Hence the auto-acceleration of acrylonitrile which persists above 60°C and exhibits the same "autoacceleration index" as at lower temperatures has to be accounted for by another cause. [Pg.244]

When translational diffusion and chemical reactions are coupled, information can be obtained on the kinetic rate constants. Expressions for the autocorrelation function in the case of unimolecular and bimolecular reactions between states of different quantum yields have been obtained. In a general form, these expressions contain a large number of terms that reflect different combinations of diffusion and reaction mechanisms. [Pg.368]

Thus, bimolecular rate constant depends only on the viscosity and the temperature of the solvent. The calculated rate constants for diffusion-controlled bimolecular reactions in solution set the upper limit for such reactions. [Pg.170]

The nearly constant peroxynitrite concentration observed in neutral solution changes dramatically when [Ft] of the solution is increased. Fig. 2 compares the transient absorption of aqueous nitrate at [Ft ] = 10-7 M and [Ft] = 0.140 M. The peroxynitrite concentration drops rapidly as protonation leads to the formation of peronitrous acid (peroxynitrous acids absorbs relatively weakly around 240 nm and is not observable in Fig. 2.). In Fig. 3 the peroxynitrite concentration is represented by the transient absorption at 310 nm as a function of [it]. As expected the formation of peroxynitrous acid increases with the concentration of protons. The protonation of peroxynitrite can be viewed as a prototypical diffusion limited bimolecular reaction and thus constitutes an excellent test bed for diffusion models. [Pg.209]

The Hanusse theorem [23] discussed in Section 2.1.1 was later generalized for the case of diffusion by Tyson and Light [32], Therefore, the mono- and bimolecular reactions with one or two intermediate products are expected to strive asymptotically, as t —> oo, for the stationary spatially-homogeneous solution Ci(r, oo) = nt(oo) corresponding to equations (2.1.2) for a system with the complete particle mixing. [Pg.68]

Lastly, we would like to mention here results of the two kinds of large-scale computer simulations of diffusion-controlled bimolecular reactions [33, 48], In the former paper [48] reactions were simulated using random walks on a d-dimensional (1 to 4) hypercubic lattice with the imposed periodic boundary conditions. In the particular case of the A + B - 0 reaction, D = Dq and nA(0) = nB(0), the critical exponents 0.26 0.01 0.50 0.02 and 0.89 0.02 were obtained for d = 1 to 3 respectively. The theoretical value of a = 0.75 expected for d = 3 was not achieved due to cluster size effects. The result for d = 4, a = 1.02 0.02, confirms that this is a marginal dimension. However, in the case of the A + B — B reaction with DB = 0, the asymptotic longtime behaviour, equation (2.1.106), was not achieved at all - even at very long reaction times of 105 Monte Carlo steps, which were sufficient for all other kinds of bimolecular reactions simulated. It was concluded that in practice this theoretically derived asymptotics is hardly accessible. [Pg.353]

Although the extinction coefficient of the benzyl radical was not known, product analysis gave an approximate total yield of benzyl radicals, from whence it could be deduced that e318 a 1,100 l.mole-1.cm-1 and the rate coefficients for the second order processes were 4xl07 and 2x 108 l.mole-1.sec 1 respectively. These rate coefficients are almost equal to those for the diffusion-controlled bimolecular reaction. [Pg.90]

Shoup, G. Lipari, and A. Szabo, Diffusion-controlled bimolecular reaction rates- the effect of rotational diffusion and orientation, Biophys. J. 36, 697-714 (1981). [Pg.116]

In this chapter we consider chemical reactions in solution first, how solvents modify the potential energy surface of the reacting molecules, and second the role of diffusion. The reactants of bimolecular reactions are brought into contact by diffusion, and there will therefore be an interplay between diffusion and chemical reaction that determines the overall reaction rate. The results are as follows. [Pg.223]

Chapters 9-11 deal with elementary reactions in condensed phases. Chapter 9 is on the energetics of solvation and, for bimolecular reactions, the important interplay between diffusion and chemical reaction. Chapter 10 is on the calculation of reaction rates according to transition-state theory, including static solvent effects that are taken into account via the so-called potential-of-mean force. Finally, in Chapter 11, we describe how dynamical effects of the solvent may influence the rate constant, starting with Kramers theory and continuing with the more recent Grote-Hynes theory for... [Pg.385]

The estimated values of the p s and of the original parameters are shown in Table V. Despite the wide range of initiator and monomer concentrations used, it is not possible to obtain precise estimates of this many parameters from the data. In particular, ps is very poorly defined for this system. Notice that the geometric mean approximation is equivalent to fis = 1 (see Equation 15). For a diffusion-controlled bimolecular reaction the arithmetic mean is appropriate as shown above, and this is reflected in the fact that p3 is significantly less than 1. [Pg.57]

A.2.2. Diffusion-Controlled Rate Constant Recently, we have calculated the diffusion-controlled (i.e., attainable maximum) rate constant of ET at an OAV interface [49]. Figure 8.8 shows models for diffusion-controlled bimolecular reactions (a) in homogeneous solution and (b) at an O/W interface. [Pg.181]

If we compare Eq. (XV.2.8) with Eq. (XV.2.3), we see that the latter is about twice as large. This is to be expected because the latter measures the frequency of all A-B encounters, while Eq. (XV.2.8) measures only new encounters. Collins and KimbalP have pointed out that in a diffusion-controlled bimolecular reaction between A and B, the initial rate which can be characterized by a random spatial distribution of A and B decays to the lower rate given by Eq. (XV.2.9). The reason for this is that the reaction tends to draw off the A-B pairs in close proximity and leaves a stationary distribution of A-B which approaches that given by the concentration gradient of Eq. (XV.2.6). The relaxation time for such a decay is of the order of " riB/ir AB, which for most molecular systems will be of the order of 10 sec, or the actual time of an encounter. Noyes has shown that there exist certain experimental systems in which these effects can be observed. We shall say more about them later in our discussion of cage effects in liquids. [Pg.498]

Abstract A challenging task in surface science is to unravel the dynamics of molecules on surfaces associated with, for example, surface molecular motion and (bimolecular) reactions. As these processes typically take place on femtosecond time scales, ultrafast lasers must be used in these studies. We demonstrate two complementary approaches to study these ultrafast molecular dynamics at metal surfaces. In the first, the molecules are studied after desorbing from the surface initiated by a laser pulse using the so called time-of-flight technique. In the second approach, molecules are studied in real time during their diffusion over the surface by using surface-specific pump-probe spectroscopy. [Pg.203]

Northrup, S. H., S. A. Allison and J. A. McCammon. (1984). Brownian dynamics simulations of diffusion influenced bimolecular reactions. J Chem. Phys. 80 1517. [Pg.232]

Transport is an integral component of all reaction systems. In well-mixed homogeneous solutions, the concentrations of all reactants and products are the same throughout the system, and there is no net movement of chemicals in space. The role of mass transport becomes evident only when chemical reactions are extremely fast. Diffusion determines the encounter frequency of reacting molecules and sets an upward limit on overall rates of reaction. (For example, for a diffusion-controlled bimolecular reaction in water the reaction rate constant is on the order of 1010 to 1011 M 1 s"1.) Mass transport plays a pronounced role in surface chemical reactions, since net movement of reactants (from solution to the surface) and products (from the surface to solution) often takes place. [Pg.25]

Note that r and the diffusion coefficient D have cancelled from Equation 2.29, because D is inversely proportional to the molecular radii r /2. Hence the rate constant kd depends only on temperature and solvent viscosity in this approximation. A selection of viscosities of common solvents and rate constants of diffusion as calculated by Equation 2.29 is given in Table 8.3. The effect of diffusion on bimolecular reaction rates is often studied by changing either the temperature or the solvent composition at a given temperature. For many solvents,54-56 although not for alcohols,57 the dependence of viscosity on temperature obeys an Arrhenius equation, that is, plots of log rj versus 1 IT are linear over a considerable range of temperatures and so are plots of log(kdr]/T) versus 1/T.56... [Pg.46]

Experiments carried out at low temperature are complimented by flash photolysis studies performed at room temperature. At low temperature, particularly in rigid media, reactive intermediates are stabilized because the rates of their unimolecular reactions are slowed, and bimolecular reactions are prevented by inhibition of diffusion. As we have just seen, this increased stability enables the application of a variety of spectroscopic methods which can aid in the determination of the structure of the intermediates. Flash photolysis experiments permit the study of absolute reactivity. These experiments can be carried out in the very short time scale required to monitor progress of reactive intermediates to stable products. In principle, the dual approach should permit thorough characterization low temperature methods reveal structure, flash photolysis probes reactivity. In practice, and particularly for the case of the aryl azides, complications can arise when the... [Pg.103]

Most of the theory of diffusion and chemical reaction in gas-solid catalytic systems has been developed for these simple, unimolecular and irreversible reactions (SUIR). Of course this is understandable due to the obvious simplicity associated with this simple network both conceptually and practically. However, most industrial reactions are more complex than this SUIR, and this complexity varies considerably from single irreversible but bimolecular reactions to multiple reversible multimolecular reactions. For single reactions which are bimolecular but still irreversible, one of the added complexities associated with this case is the non-monotonic kinetics which lead to bifurcation (multiplicity) behaviour even under isothermal conditions. When the diffusivities of the different components are close to each other that added complexity may be the only one. However, when the diffusiv-ities of the different components are appreciably different, then extra complexities may arise. For reversible reactions added phenomena are introduced one of them is discussed in connection with the ammonia synthesis reaction in chapter 6. [Pg.89]

The transition state theory (TST) may be considered to be established in 1941 by publication of a momunental book The Theory of Rate Processes [1. In Chapter VIII of the book, the authors discuss solution reactions and conclude. . that the ratedetermining step in solution is. .. the formation from the reactants of an activated complex which subsequently decomposes . Though the authors pointed out the importance of diffusion in bimolecular reactions, they did not consider a possible break down of their two key assumptions, that is, thermal equilibrium between the initial and the transition state and neglecting recrossing, in imimolecular rate processes. The remarkable success of TST in the interpretation of kinetic effects of pressure [2] turned the attention of high-pressure kineticists away from a possible failure of TST and efforts were concentrated on the interpretation of the activation volume obtained from pressure dependence of a rate constant fe at a constant temperature (Eq. 3.1). [Pg.97]


See other pages where Diffusion and bimolecular reactions is mentioned: [Pg.843]    [Pg.231]    [Pg.233]    [Pg.235]    [Pg.237]    [Pg.130]    [Pg.843]    [Pg.843]    [Pg.231]    [Pg.233]    [Pg.235]    [Pg.237]    [Pg.130]    [Pg.843]    [Pg.2946]    [Pg.104]    [Pg.127]    [Pg.235]    [Pg.142]    [Pg.191]    [Pg.225]    [Pg.73]    [Pg.397]    [Pg.217]    [Pg.227]    [Pg.451]    [Pg.191]   


SEARCH



Diffusion and reaction

Diffusion reactions

Diffusivity reactions

© 2024 chempedia.info