Big Chemical Encyclopedia

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

Articles Figures Tables About

Reaction schemes, kinetic analysis

We first consider the simple example of an uncatalyzed Diels-Alder reaction shown in Scheme 50.2 in order to demonstrate the use of different excess experiments. The Diels-Alder reaction is known to exhibit second overall order kinetics, as shown in eq. (4). We demonstrate with this known case how reaction progress kinetic analysis may be used to extract the reaction orders in both substrate concentrations, [5] and [6]. ... [Pg.448]

Reaction progress kinetic analysis offers a reliable alternative method to assess the stability of the active catalyst concentration, again based on our concept of excess [e]. In contrast to our different excess experiments described above, now we carry out a set of experiments at the same value of excess [ej. We consider again the proline-mediated aldol reaction shown in Scheme 50.1. Under reaction conditions, the proline catalyst can undergo side reactions with aldehydes to form inactive cyclic species called oxazolidinones, effectively decreasing the active catalyst concentration. It has recently been shown that addition of small amounts of water to the reaction mixture can eliminate this catalyst deactivation. Reaction progress kinetic analysis of experiments carried out at the same excess [e] can be used to confirm the deactivation of proline in the absence of added water as well to demonstrate that the proline concentration remains constant when water is present. [Pg.452]

The objective of this paper is to demonstrate the importance of phase and reaction equilibria considerations in the rational development of SCF reaction schemes. Theoretical analysis of phase and reaction equilibria are presented for two relatively simple reactions, viz., the isomerizations of n-hexane and 1-hexene. Our simulated conversion and yield plots compare well with experimental results reported in the literature for n-hexane isomerization (4) and obtained by us for 1-hexene isomerization. Based on our analysis, the choice of an appropriate SCF reaction medium for each of these reactions is discussed. Properties such as viscosity, surface tension and polarity can affect transport and kinetic behavior and hence should also be considered for complete evaluation of SCF solvents. These rate effects are not considered in our equilibrium study. [Pg.302]

In our illustration of the graphical manipulations of data using reaction progress kinetic analysis, we will make use of the example of a model reaction, the intermolecnlar aldol reaction between acetone 1 and aldehyde 2 to form the aldol addition product 3, mediated by proline 4, as shown in Scheme 27.1. The demonstration by List, Lemer, and Barbas in 2000 that proline mediates intermolecular aldol reactions with a high degree of asymmetric induction heralded a revolution in the field of organocatalysis, encompassing the discovery of new catalysts and new catalytic transformations." ... [Pg.457]

The key to reaction progress kinetic analysis lies in making use of the reaction stoichiometry. For the aldol reaction of Scheme 27.1, a mass balance tells us that for each molecule of ketone 1 consumed, one molecule of aldehyde 2 is also consumed. The parameter of interest for our analysis is the difference between the initial concentrations of the two substrates. This parameter is a constant in any given experiment, and we call it the excess, or [e], with units of molarity (Equation 27.5). [Pg.458]

Now we are prepared to illustrate these experimental protocols of reaction progress kinetic analysis using data from reaction calorimetric monitoring of the aldol reaction shown in Scheme 27.1. We turn hrst to the issue of catalyst stability using our same excess protocol. In these aldol reactions, it was noted that the active catalyst concentration can be effectively decreased by the formation of oxazolidinones between proline and aldehydes or ketones, and that addition of water can suppress this catalyst deactivation. Same excess reactions carried out in the absence of water and in the presence of water are shown in Figure 27.3a and Figure 27.3b, respectively. The plots do not overlay in the absence of water, but they do when water is present. The overlay in these same [e] experiments in Figure 27.3b means that the total concentration of active catalyst within the cycle is constant and is the same in the two experiments where water is present. [Pg.461]

This chapter describes the effects of micelles on the Diels-Alder reaction of compounds 5,1 a-g (see Scheme 5.1) with cyclopentadiene (5.2). As far as we know, our study is the first detailed kinetic analysis of micellar catalysis of a Diels-Alder reaction. [Pg.132]

A useful approach that is often used in analysis and simplification of kinetic expressions is the steady-state approximation. It can be illustrated with a hypothetical reaction scheme ... [Pg.195]

The method of evaluation of the rate constants for this reaction scheme will depend upon the type of analytical information available. This depends in part upon the nature of the reaction, but it also depends upon the contemporary state of analytical chemistry. Up to the middle of the 20th century, titrimetry was a widely applied means of studying reaction kinetics. Titrimetric analysis is not highly sensitive, nor is it very selective, but it is accurate and has the considerable advantage of providing absolute concentrations. When used to study the A —> B — C system in which the same substance is either produced or consumed in each step (e.g., the hydrolysis of a diamide or a diester), titration results yield a quantity F = Cb + 2cc- Swain devised a technique, called the time-ratio method, to evaluate the rate... [Pg.69]

However, there is an important difference between these two systems in the ligand-metal ion ratio in complexation. Namely, micellar reactions require a more generalized reaction Scheme 3, where the molarity of ligand n is either 1 or 2 depending upon the structure of the ligands. This scheme gives rates Eq. 2-4 for n = 1 and Eq. 3, 5, 6 for n = 2. The results of the kinetic analysis are shown in Table 3. [Pg.156]

For most real systems, particularly those in solution, we must settle for less. The kinetic analysis will reveal the number of transition states. That is, from the rate equation one can count the number of elementary reactions participating in the reaction, discounting any very fast ones that may be needed for mass balance but not for the kinetic data. Each step in the reaction has its own transition state. The kinetic scheme will show whether these transition states occur in succession or in parallel and whether kinetically significant reaction intermediates arise at any stage. For a multistep process one sometimes refers to the transition state. Here the allusion is to the transition state for the rate-controlling step. [Pg.126]

A reader familiar with the first edition will be able to see that the second derives from it. The objective of this edition remains the same to present those aspects of chemical kinetics that will aid scientists who are interested in characterizing the mechanisms of chemical reactions. The additions and changes have been quite substantial. The differences lie in the extent and thoroughness of the treatments given, the expansion to include new reaction schemes, the more detailed treatment of complex kinetic schemes, the analysis of steady-state and other approximations, the study of reaction intermediates, and the introduction of numerical solutions for complex patterns. [Pg.293]

Toward these ends, the kinetics of a wider set of reaction schemes is presented in the text, to make the solutions available for convenient reference. The steady-state approach is covered more extensively, and the mathematics of other approximations ( improved steady-state and prior-equilibrium) is given and compared. Coverage of data analysis and curve fitting has been greatly expanded, with an emphasis on nonlinear least-squares regression. [Pg.293]

Kinetic Analysis. The following reaction scheme is proposed to account for the observed ionic yields shown in Figure 10 in a system primarily composed of nitrogen and water vapor. [Pg.336]

Cortright RD, Dumesic JA. 2001. Kinetics of heterogeneous catal)4ic reactions analysis of reaction schemes. Adv Catal 46 161 -264. [Pg.88]

Reaction step 5 in Scheme 3.1 can be rnled ont becanse the flnoranil ketyl radical (FAH ) reaches a maximum concentration within 100 ns as the triplet state ( FA) decays by reaction step 2 while the fluoranil radical anion (FA ) takes more than 500 ns to reach a maximum concentration. This difference snggests that the flnoranil radical anion (FA ) is being produced from the fluoranil ketyl radical (FAH ). Reaction steps 1 and 2 are the most likely pathway for prodncing the flnoranil ketyl radical (FAH ) from the triplet state ( FA) and is consistent with the TR resnlts above and other experiments in the literatnre. The kinetic analysis of the TR experiments indicates the fluoranil radical anion (FA ) is being prodnced with a hrst order rate constant and not a second order rate constant. This can be nsed to rnle ont reaction step 4 and indicates that the flnoranil radical anion (FA ) is being prodnced by reaction step 3. Therefore, the reaction mechanism for the intermolecular hydrogen abstraction reaction of fluoranil with 2-propanol is likely to predominantly occur through reaction steps 1 to 3. [Pg.155]

We emphasize that the critical ion pair stilbene+, CA in the two photoactivation methodologies (i.e., charge-transfer activation as well as chloranil activation) is the same, and the different multiplicities of the ion pairs control only the timescale of reaction sequences.14 Moreover, based on the detailed kinetic analysis of the time-resolved absorption spectra and the effect of solvent polarity (and added salt) on photochemical efficiencies for the oxetane formation, it is readily concluded that the initially formed ion pair undergoes a slow coupling (kc - 108 s-1). Thus competition to form solvent-separated ion pairs as well as back electron transfer limits the quantum yields of oxetane production. Such ion-pair dynamics are readily modulated by choosing a solvent of low polarity for the efficient production of oxetane. Also note that a similar electron-transfer mechanism was demonstrated for the cycloaddition of a variety of diarylacetylenes with a quinone via the [D, A] complex56 (Scheme 12). [Pg.217]

In eq. 8 are shown the results of a kinetic analysis of the series of reactions in Scheme 1. The analysis is based on the quenching rate constant k, corrected for diffusional effects, which would be measured for the quenching of Ru(bpy>3 + by PQ +. [Pg.157]

It appears like a miracle how aliphatic chains (mainly olefins and paraffins) are formed from a mixture of CO and H2. But miracle means only high complexity of unknown order (Figure 9.1). Problems in FT synthesis research include the visualization of a multistep reaction scheme where adsorbed intermediates are not easily identified. Kinetic constants of the elemental reactions are not directly accessible. Models and assumptions are needed. The steady state develops slowly. The true catalyst is assembled under reaction conditions. Difficulties with product analysis result from the presence of hundreds of compounds (gases, liquids, solids) and from changes of composition with time. [Pg.166]

It should be emphasized that clear-cut situations described in Schemes 1-3 are uncommon and typically the combination of these models needs to be considered for kinetic and mechanistic description of a real system. However, even when one of the limiting cases prevails, each of these models may predict very different formal kinetic patterns depending on where the rate determining step is located. For the same reason, different schemes may be consistent with the same experimental rate law, i.e. thorough formal kinetic description of a reaction and the analysis of the rate law may not be conclusive with respect to the mechanism of the autoxidation process. [Pg.399]

If electron transport is fast, the system passes from zone R to zone S+R and then to zone SR. In the latter case there is a mutual compensation of diffusion and chemical reaction, making the substrate concentration profile decrease within a thin reaction layer adjacent to the film-solution interface. This situation is similar to what we have termed pure kinetic conditions in the analysis of an EC reaction scheme adjacent to the electrode solution interface developed in Section 2.2.1. From there, if electron transport starts to interfere, one passes from zone SR to zone SR+E and ultimately to zone E, where the response is controlled entirely by electron transport. [Pg.290]

The chlorine atom adds in the gas phase to propadiene (la) with a rate constant that is close to the gas-kinetic limit. According to the data from laser flash photolysis experiments, this step furnishes exclusively the 2-chloroallyl radical (2a) [16, 36], A computational analysis of this reaction indicates that the chlorine atom encounters no detectable energy barrier as it adds either to Ca or to Cp in diene la to furnish chlorinated radical 2a or 3a. A comparison between experimental and computed heats of formation points to a significant thermochemical preference for 2-chloroal-lyl radical formation in this reaction (Scheme 11.2). Due to the exothermicity of both addition steps, intermediates 2a and 3a are formed with considerable excess energy, thus allowing isomerizations of the primary adducts to follow. [Pg.704]

In order to investigate in more depth the mechanism of Scheme 4.10, a detailed kinetic analysis has been performed by choosing the methylation of phe-nylacetonitrile (2a) and methyl phenylacetate (2b) with DMC as model reactions. Some general considerations are the following. [Pg.89]

The bromination reaction (Scheme 12.3) was also carried out on resins (1) of three different sizes (Fig. 12.5). Single bead FTIR study and the kinetics analysis were carried out as in the esterification reaction studies. Rate constants are hsted in Tab. 12.2. The relationship between the rate constants and the bead size is shown in Fig. 12.7b. [Pg.510]

In a dehydration reaction (Scheme 12.4), the IR band of the formamide carbonyl group at 1684 cm in (7) decreased and eventually converted to the isonitrile band at 2150 cm in (8) (Fig. 12.8). In a separate example (Scheme 12.5), the conversion of the IR band from the carbonate carbonyl group in (9) to the IR band of the carbamide carbonyl group in (10) can be monitored to assure the reaction completion (Fig. 12.9). Based on FTIR analysis, the reaction time course can be analyzed by integrating peak areas of the IR bands from the starting resin and the product. From the point of view of kinetics, the side reaction product formation can be excluded if the pseudo first order rates of the starting material consumption and the product formation are identical. [Pg.510]

Rate constants for the reaction of thiyl radicals with the t-BuMePhSiH were also extracted from the kinetic analysis of the thiol-catalysed radical-chain racemization of enantiomerically pure (S)-isomer [34]. Scheme 3.2 shows the reaction mechanism that involves the rapid inversion of silyl radicals together with reactions of interest. The values in cyclohexane solvent at 60 °C are collected in the last column of Table 3.5. [Pg.42]

The chain ion-radical mechanism of ter Meer reaction has been supported by a thorough kinetic analysis. The reaction is well-described by a standard equation of chain-radical processes (with square-law chain termination) (Shugalei et al. 1981). This mechanism also explains the nature of side products—aldehydes (see steps 13 and 14) as well as vicinal dinitroethylenes. Scheme 4.37 explains formation of vic-dinitroethylenes. [Pg.245]

The formation of l,4-dihydro-2,3-benzodioxin 5 from the benzocyclobutene 141/o-quinodimethane 142 equilibrium has been utilized as a trapping experiment for the kinetic analysis of diradical reactions (Scheme 38) <2002CC1594, 1988CB1357>. [Pg.712]

Under these conditions the increase in G(CH3COCH3) should be paralleled by an increase in G(H202), as observed. If Reactions 8 and 9 are rapid relative to Reaction 18, then it follows from kinetic analysis of the above reaction scheme, assuming a stationary concentration of H02 radicals, that the yield of acetone arising from the chain reaction is given by... [Pg.118]

A review of the Journal of Physical Chemistry A, volume 110, issues 6 and 7, reveals that computational chemistry plays a major or supporting role in the majority of papers. Computational tools include use of large Gaussian basis sets and density functional theory, molecular mechanics, and molecular dynamics. There were quantum chemistry studies of complex reaction schemes to create detailed reaction potential energy surfaces/maps, molecular mechanics and molecular dynamics studies of larger chemical systems, and conformational analysis studies. Spectroscopic methods included photoelectron spectroscopy, microwave spectroscopy circular dichroism, IR, UV-vis, EPR, ENDOR, and ENDOR induced EPR. The kinetics papers focused on elucidation of complex mechanisms and potential energy reaction coordinate surfaces. [Pg.178]

For the most reactive compounds in the series, i.e., 4,6-dinitro-benzofurazans and 4,6-dinitrobenzofuroxans, not only water222 but also methanol220 is found to be an effective neutral nucleophile in the appropriate pH range. The thermodynamic and kinetic analysis for the reactions in methanol similar to that described above for the reactions in water has been applied to the formation of 172 from 4,6-dinitro-7-methoxybenzofurazan (195) in the pH range from 2.2 to 14. The formation of 172 has been followed spectrophotometrically and found to be complete in methanol. The reaction scheme is as follows... [Pg.434]

A more conclusive kinetic analysis is not only made difficult because of the temperature-dependent occurrence of different phosphoprotein species, but also by the fact that the spontaneous decay of gradient-independent as well as of gradient-dependent phosphoprotein depends on the concentration of phosphate. This phosphate dependence is not taken into account by the tentative reaction scheme. [Pg.50]

The complicated reaction scheme for the dehydration of alcohols (Scheme 2) makes kinetic analysis rather difficult. However, initial reaction rates have been measured, without special problems, for secondary... [Pg.282]


See other pages where Reaction schemes, kinetic analysis is mentioned: [Pg.447]    [Pg.1263]    [Pg.295]    [Pg.17]    [Pg.469]    [Pg.186]    [Pg.352]    [Pg.104]    [Pg.84]    [Pg.189]    [Pg.236]    [Pg.500]    [Pg.218]    [Pg.953]    [Pg.69]   
See also in sourсe #XX -- [ Pg.36 ]




SEARCH



Analysis, scheme

Kinetic analysis

Kinetic scheme

Kinetics scheme

Reaction scheme

© 2024 chempedia.info