Big Chemical Encyclopedia

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

Articles Figures Tables About

Kinetics at High Conversions

If crosslinking or copolymer precipitation occurs, bulk polymerization may be difficult to handle. A suspension process may then be the only feasible way in which the copolymerization can be carried out [4]. Suspension processes also provide a means of investigating copolymerization kinetics at high conversion. The monomer sequence in styrene-methyl methacrylate copolymers at high conversion have been found to differ from those observed at low conversion [77]. [Pg.230]

O Shaughnessy B, Yn J. Non-steady state free radical polymerization kinetics at high conversions entangled regimes. Macromolecnles 1998 31 5240-5354. [Pg.225]

Initially, it was thought more likely that the electron poor metal atom would be involved in the electrophilic attack at the alkene and also the metal-carbon bond would bring the alkene closer to the chiral metal-ligand environment. This mechanism is analogous to alkene metathesis in which a metallacyclobutane is formed. Later work, though, has shown that for osmium the actual mechanism is the 3+2 addition. Molecular modelling lends support to the 3+2 mechanism, but also kinetic isotope effects support this (KIEs for 13C in substrate at high conversion). Oxetane formation should lead to a different KIE for the two alkene carbon atoms involved. Both experimentally and theoretically an equal KIE was found for both carbon atoms and thus it was concluded that an effectively symmetric addition, such as the 3+2 addition, is the actual mechanism [22] for osmium. [Pg.312]

As Table 5 shows, the volume of a continuous stirred tank with a certain performance is greater than that of the corresponding plug-flow reactor. The volume ratio with a second-order reaction is markedly greater than when first-order kinetics apply and this effect is greater at high conversions where both ratios can be very large. [Pg.91]

Thus it is evident that a PFTR is always the reactor of choice (smaller for greater than zero-order kinetics in an isothermal reactor. The CSTR may stUl be favored for n > 0 for cost reasons as long as the conversion is not too high, but the isothermal PFTR is much superior at high conversions whenever n > 0. [Pg.100]

In the limiting cases (6) and (7), simple first order kinetics is expected. In the intermediate case (5), the kinetics is complex as seen in Fig. 4. At low contact time and conversion, r2 is negligible and the slope of the conversion curve represents k- at high conversion, the slope reflects k2. In this regime it is not admissible to do a single experiment and from the observed conversion calculate a rate constant by assuming first order kinetics. Apparent "rate constants osculated in this way yield variable values that depend on the arbitrarily chosen conversion (slope of dashed lines, Fig. 4). [Pg.261]

Hoft reported about the kinetic resolution of THPO (16b) by acylation catalyzed by different lipases (equation 12) °. Using lipases from Pseudomonas fluorescens, only low ee values were obtained even at high conversions of the hydroperoxide (best result after 96 hours with lipase PS conversion of 83% and ee of 37%). Better results were achieved by the same authors using pancreatin as a catalyst. With this lipase an ee of 96% could be obtained but only at high conversions (85%), so that the enantiomerically enriched (5 )-16b was isolated in poor yields (<20%). Unfortunately, this procedure was limited to secondary hydroperoxides. With tertiary 1-methyl-1-phenylpropyl hydroperoxide (17a) or 1-cyclohexyl-1-phenylethyl hydroperoxide (17b) no reaction was observed. The kinetic resolution of racemic hydroperoxides can also be achieved by chloroperoxidase (CPO) or Coprinus peroxidase (CiP) catalyzed enantioselective sulfoxidation of prochiral sulfides 22 with a racemic mixmre of chiral hydroperoxides. In 1992, Wong and coworkers and later Hoft and coworkers in 1995 ° investigated the CPO-catalyzed sulfoxidation with several chiral racemic hydroperoxides while the CiP-catalyzed kinetic resolution of phenylethyl hydroperoxide 16a was reported by Adam and coworkers (equation 13). The results are summarized in Table 4. [Pg.332]

In order to quantify diffiisional effects on curing reactions, kinetic models are proposed in the literature [7,54,88,95,99,127-133]. Special techniques, such as dielectric permittivity, dielectric loss factor, ionic conductivity, and dipole relaxation time, are employed because spectroscopic techniques (e.g., FT i.r. or n.m.r.) are ineffective because of the insolubility of the reaction mixture at high conversions. A simple model, Equation 2.23, is presented by Chem and Poehlein [3], where a diffiisional factor,//, is introduced in the phenomenological equation, Equation 2.1. [Pg.84]

However, there are some contradictory reports on the composition of the products of toluene alkylation or benzene dialkylation at high conversions. In some cases, compositions corresponding to the thermodynamic equilibrium between ortho, meta and para isomers were found, and in other cases, kinetic control of orientation, giving mostly the ortho + para substitution, prevailed. Consecutive isomerisation of the ortho and para isomers to the more stable meta isomer seems to be the cause of the disagreement. More active catalysts gave more meta derivatives than the less active ones [343] and increasing the temperature has the same effect [351]. [Pg.336]

We begin by describing the current understanding of the kinetics of polymerization of classical unsaturated monomers and macromonomers in the disperse systems. In particular, we note the importance of diffusion-controlled reactions of such monomers at high conversions, the nucleation mechanism of particle formation, and the kinetics and kinetic models for radical polymerization in disperse systems. [Pg.7]

Differences in product yields can be amplified when low conversions are screened because the initial kinetics are monitored rather than integral productivities at high conversions. This can strategically be used to facilitate catalyst comparisons. [Pg.72]

In the example shown in Figure 7.5, a 25% molar excess of B has been added (M = 1.25). The function of this excess is to increase the reaction rate when CA becomes small. It is often called the kinetic excess, because it increases the reaction rate at high conversion (Equation 7.7). [Pg.157]

Reactive extrusion has emerged from a scientific curiosity to an industrial process. Various types of extruders can be used, all with their specific advantages and disadvantages. Further development suffers from lack of kinetic and rheological data at high conversions and from uncertainties about heat transfer and reactor stability. Nonlinear effects in the process can give rise to instabilities that are of thermal, hydrodynamical or chemical origin. [Pg.393]

Xie, T.Y, Hamielec, A.E., Wood P.E., Woods, D.R., Experimental investigation of vinyl chloride polymerisation at high conversion mechanism, kinetics and modeling, Polymer, 1991, 32(6), 537-557... [Pg.396]

The rational design of a reaction system to produce a desired polymer is more feasible today by virtue of mathematical tools which permit one to predict product distribution as affected by reactor type and conditions. New analytical tools such as gel permeation chromatography are beginning to be used to check technical predictions and to aid in defining molecular parameters as they affect product properties. The vast majority of work concerns bulk or solution polymerization in isothermal batch or continuous stirred tank reactors. There is a clear need to develop techniques to permit fuller application of reaction engineering to realistic nonisothermal systems, emulsion systems, and systems at high conversion found industrially. A mathematical framework is also needed which will start with carefully planned experimental data and efficiently indicate a polymerization mechanism and statistical estimates of kinetic constants rather than vice-versa. [Pg.18]

The Kinetics of Methanol Carbonylation Over RhX, RhY and IrY zeolites Carbonylation of methanol proceeds readily at atmospheric pressure under mild temperature conditions 150°-180°C. This reaction ZCH OH + CO - CH COOCH + HjO produces mainly methyl acetate and water. Acetic acid was detected at high conversions and high temperatures. Traces of dimethyl ether could also form. In most cases the selectivity to methyl acetate was at least 90% in presence of the iodide promotor. [Pg.462]

Rate constants change in the course of MMA and VAC polymerizations [39, 40] (Tables 3 and 4). The considerable decrease in k, at high conversion is caused by a switch of the rate-determining step from the kinetic domain to a diffusion-controlled process [41] (see Chap. 6, Sect. 3.1)... [Pg.173]


See other pages where Kinetics at High Conversions is mentioned: [Pg.215]    [Pg.98]    [Pg.66]    [Pg.102]    [Pg.715]    [Pg.1223]    [Pg.215]    [Pg.98]    [Pg.66]    [Pg.102]    [Pg.715]    [Pg.1223]    [Pg.126]    [Pg.227]    [Pg.81]    [Pg.202]    [Pg.203]    [Pg.275]    [Pg.285]    [Pg.248]    [Pg.48]    [Pg.72]    [Pg.58]    [Pg.555]    [Pg.228]    [Pg.332]    [Pg.555]    [Pg.369]    [Pg.126]    [Pg.201]    [Pg.21]    [Pg.195]    [Pg.298]    [Pg.46]    [Pg.191]    [Pg.196]    [Pg.423]    [Pg.196]    [Pg.71]   


SEARCH



Conversion kinetics

© 2024 chempedia.info