Big Chemical Encyclopedia

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

Articles Figures Tables About

Reaction predicting product when both

In hindsight, the primary factor in determining which approach is most applicable to a particular reacting flow is the characteristic time scales of the chemical reactions relative to the turbulence time scales. In the early applications of the CRE approach, the chemical time scales were larger than the turbulence time scales. In this case, one can safely ignore the details of the flow. Likewise, in early applications of the FM approach to combustion, all chemical time scales were assumed to be much smaller than the turbulence time scales. In this case, the details of the chemical kinetics are of no importance, and one is free to concentrate on how the heat released by the reactions interacts with the turbulent flow. More recently, the shortcomings of each of these approaches have become apparent when applied to systems wherein some of the chemical time scales overlap with the turbulence time scales. In this case, an accurate description of both the turbulent flow and the chemistry is required to predict product yields and selectivities accurately. [Pg.21]

The photocyclodehydrogenation of thienyl ethylenes is well-defined when both thiophene rings are bound via a C(2) atom to the ethylenic bond as in (70). In other cases, however, more cyclization products are possible. To predict the photocyclization mode for heterohelicenes the F s rule fails in many cases, because correction factors for the hetero atoms in the Huckel MO calculation have to be introduced and the systems are not well comparable with carbocyclic diaryl ethylenes. A better reaction parameter in these cases is the Mulliken overlap population (nrs)51), introduced by Muszkat52) for these cases. The overlap populations of the atoms r and s in ground and excited state (nIiS and n s), are calculated using the extended Huckel method. Cyclizations should not occur when nr>s and An s (= nr>s — n s) have negative values. (This method can also be used for diaryl olefins, but in these cases calculation of F s is more simple.). [Pg.78]

In a Diels-Alder reaction, when both n systems are polarized, the more favorable overlap, and therefore the stronger interaction, occurs when the ends with the larger coefficients get together and the smaller coefficients get together. Predict the major product of each of the following cycloaddition reactions ... [Pg.288]

There is also a need for chemical reaction engineering courses to deal more thoroughly with the chemistry of the process under consideration. This is particularly important when both product quality and yield are the performance targets. The use of modem concepts of physical chemistry to make predictions of transport and rate parameters should also be emphasized, since such concepts show how the properties of a system affect these parameters. [Pg.224]

Stereochemistry of CO Dissociation. If a complex has nonequivalent CO groups, the stereochemistry of the dissociative process can be defined. This has prompted many studies of exchange reactions with Mn(CO)5Br. Substitution reactions always lead to m-Mn(CO)4LBr, but substitution products do not necessarily provide information on the site of CO dissociation. In an exchange reaction, the products must reflect the site of dissociation by the principle of microscopic reversibility. One would predict that the site of CO dissociation would be cis to the bromide in Mn(CO)sBr since each cis-CO (equatorial) is trans to CO and should not be as strongly jt bonded as the axial CO, which is trans to Br. However, both the axial and equatorial positions were emiched when reacted withC Oor CO ... [Pg.2568]

The issue of simple diastereoselectivity arises when both the allyl and vinyl moieties of the ketene N,0-acetal intermediate 2 are substituted at their terminus, leading to vicinal stereocenters in the products (Scheme 7.22). In analogy to the aldol reaction, the stereochemical outcome can be predicted in terms of a Zimmermann-Traxler type chair-shaped transition state. Accordingly, the synlanti ratio of the products depends on double bond geometry. Whereas the geometry of the unsaturated alcohol is pre-determined and usually not subject to equilibration, the geometry of the ketene N,0-acetal moiety depends on the reaction conditions that lead to its in situ formation. [Pg.382]

When both 1,3-dipole and dipolarophile are unsymmetrical, two products are possible. The formation of major product can be predicted by consideration of their TSs. The most stable TS will provide the major product. The stabihty of the TS is controlled by both electronic and steric factors. Therefore, the regioselectivity of a 1,3-DPCA reaction is determined by the steric and electronic properties of the substituents attached to 1,3-dipole and dipolarophile. The FMO theory may also be applied to analyze the regioselectivity of 1,3-DPCA reaction [107]. A relatively stronger donor-acceptor interaction between HOMO and LUMO and lowest dipole moment favors the TS. The HOMO and LUMO of a 1,3-dipole are similar to that of a diene in a Diels-Alder reaction. The interactions of HOMO or LUMO of a dipole with a LUMO or HOMO of a dipolarophile depend on their electron donor and electron acceptor property. The orbital interactions of HOMO and LUMO of dipole and dipolarophile are shown in Fig. 3.11. [Pg.81]

There is an additional problem in [2 + 4] cyclo-additions concerning the relative orientations of the two components when both are unsymmetrical. Orientation effects are almost always observed in Diels-Alder reactions, and there exists a simple method for predicting the outcome of a given reaction. Consider, for example, the reactions of the dienes (40) and (41) with the olefin (42). The various possible transition state orientations are (43) and (44) for diene (40), and (45) and (46) for diene (41). The ortho and para orientations (44) or (45) are strongly preferred, and the major products are derived from... [Pg.167]

Only one exception to the clean production of two monomer molecules from the pyrolysis of dimer has been noted. When a-hydroxydi-Zvxyljlene (9) is subjected to the Gorham process, no polymer is formed, and the 16-carbon aldehyde (10) is the principal product in its stead, isolated in greater than 90% yield. This transformation indicates that, at least in this case, the cleavage of dimer proceeds in stepwise fashion rather than by a concerted process in which both methylene—methylene bonds are broken at the same time. This is consistent with the predictions of Woodward and Hoffmann from orbital symmetry considerations for such [6 + 6] cycloreversion reactions in the ground state (5). [Pg.428]

Figure 10-4 shows the predicted X as a function of A j for the 30-1 reactor at 100 rpm. Decreasing A j slows the first reaction and increases the formation of the secondary product S. As a result, the predicted Xg decreases with increasing A j. It was found that A j = 0.08 gave the best predictions when compared to the experimental data from Middleton et al. [4] Figure 10-5 shows a comparison between the experimental data from Middleton et al. and the current model predictions for both the 30-1 and 600-1 reactors. Xg is plotted as a function... [Pg.798]

Bhawe (14) has simulated the periodic operation of a photo-chemically induced free-radical polymerization which has both monomer and solvent transfer steps and a recombination termination reaction. An increase of 50% in the value of Dp was observed over and above the expected value of 2.0. An interesting feature of this work is that when very short period oscillations were employed, virtually time-invariant products were predicted. [Pg.256]

In fused ring systems, the positions are not equivalent and there is usually a preferred orientation even in the unsubstituted hydrocarbon. The preferred positions may often by predicted as for benzene rings. Thus it is possible to draw more canonical forms for the arenium ion when naphthalene is attacked at the a position than when it is attacked at the p position, and the a position is the preferred site of attack,though, as previously mentioned (p. 682), the isomer formed by substitution at the p position is thermodynamically more stable and is the product if the reaction is reversible and equilibrium is reached. Because of the more extensive delocalization of charges in the corresponding arenium ions, naphthalene is more reactive than benzene and substitution is faster at both positions. Similarly, anthracene, phenanthrene, and other fused polycyclic aromatic hydrocarbons are also substituted faster than benzene. [Pg.688]

If the carbanion has even a short lifetime, 6 and 7 will assume the most favorable conformation before the attack of W. This is of course the same for both, and when W attacks, the same product will result from each. This will be one of two possible diastereomers, so the reaction will be stereoselective but since the cis and trans isomers do not give rise to different isomers, it will not be stereospecific. Unfortunately, this prediction has not been tested on open-chain alkenes. Except for Michael-type substrates, the stereochemistry of nucleophilic addition to double bonds has been studied only in cyclic systems, where only the cis isomer exists. In these cases, the reaction has been shown to be stereoselective with syn addition reported in some cases and anti addition in others." When the reaction is performed on a Michael-type substrate, C=C—Z, the hydrogen does not arrive at the carbon directly but only through a tautomeric equilibrium. The product naturally assumes the most thermodynamically stable configuration, without relation to the direction of original attack of Y. In one such case (the addition of EtOD and of Me3CSD to tra -MeCH=CHCOOEt) predominant anti addition was found there is evidence that the stereoselectivity here results from the final protonation of the enolate, and not from the initial attack. For obvious reasons, additions to triple bonds cannot be stereospecific. As with electrophilic additions, nucleophilic additions to triple bonds are usually stereoselective and anti, though syn addition and nonstereoselective addition have also been reported. [Pg.977]


See other pages where Reaction predicting product when both is mentioned: [Pg.31]    [Pg.404]    [Pg.198]    [Pg.924]    [Pg.182]    [Pg.206]    [Pg.811]    [Pg.414]    [Pg.575]    [Pg.445]    [Pg.36]    [Pg.200]    [Pg.371]    [Pg.978]    [Pg.426]    [Pg.445]    [Pg.320]    [Pg.219]    [Pg.224]    [Pg.278]    [Pg.317]    [Pg.340]    [Pg.114]    [Pg.390]    [Pg.529]    [Pg.28]    [Pg.51]    [Pg.190]    [Pg.478]    [Pg.44]    [Pg.46]    [Pg.830]    [Pg.242]   


SEARCH



Predicting products

Predicting products reactions

Prediction production

Product prediction

Productivity prediction

Reaction prediction

© 2024 chempedia.info