You are on page 1of 11

Ind. Eng. Chem. Res.

2000, 39, 2601-2611 2601

Reaction Kinetics and Chemical Equilibrium of Homogeneously and


Heterogeneously Catalyzed Acetic Acid Esterification with Methanol
and Methyl Acetate Hydrolysis
T. Po1 pken, L. Go1 tze,† and J. Gmehling*
Carl von OssietzkysUniversität Oldenburg, Technische Chemie, P.O. Box 2503, D-26111 Oldenburg, Germany

The reaction kinetics and chemical equilibrium of the reversible esterification of methanol with
acetic acid were investigated. This system is of major importance as a model reaction for reactive
distillation. The reaction has been catalyzed both homogeneously by acetic acid itself and
heterogeneously by an acidic ion-exchange resin (Amberlyst 15). The chemical equilibrium
composition was measured for various temperatures and starting compositions of the reactants
and products. Kinetic information was obtained at temperatures between 303.15 and 343.15 K
at various starting compositions covering concentration ranges from the stoichiometric regime
to the dilute regions. Both the esterification and the hydrolysis reaction were investigated to
yield a model which is applicable for any starting composition. The homogeneous reaction has
been described with a simple power-law model. The use of activities in the kinetic model instead
of mole fractions results in a much smaller residual error. To compare pseudohomogeneous and
adsorption-based kinetic models for the heterogeneously catalyzed reaction, independent binary
liquid adsorption experiments were used to fit the adsorption constants to keep the number of
adjustable parameters the same for each model. The use of activities instead of mole fractions
results in a slight improvement of the kinetic model only, while incorporating adsorption
information into the kinetic model results in a much better fit. The chemical equilibrium
composition calculated from the kinetic model is in agreement with the measured chemical
equilibrium.

Introduction (PVA), and acetic acid and methanol can be recycled on


site into the process.
Methyl acetate synthesis by esterification of acetic
acid with methanol, and the backward reaction, the Although the reaction is known to proceed rather
hydrolysis of methyl acetate, have recently received slowly, most reactive distillation modeling and simula-
growing interest as a model reaction for reactive distil- tions considering this reaction have been performed
lation.1 Reactive distillation offers distinct advantages assuming chemical equilibrium. It has been shown by
over the conventional approach of performing the reac- Bessling et al.1 that the error introduced by assuming
tion and separation sequentially. Examples are reduced infinitely fast reaction kinetics is relatively small at
capital and operating costs due to higher conversion and atmospheric pressure and high catalyst mass. By op-
reduction of the extent of parallel and consecutive erating reactive distillation columns at reduced pressure
reactions resulting in a higher selectivity. Its major and thus reduced temperature, one is able to study the
drawback is that the chemical reaction has to have effect of reaction kinetics on the performance of a
significant conversion at distillation temperatures, which reactive distillation column and gain more insight into
limits the range of applicability somewhat. Reactive the process. This knowledge is of great importance for
distillation has been identified as being suitable for the
the design of reactive distillation columns for esterifi-
methyl acetate system for different processes, namely,
the synthesis of methyl acetate,2,3 the hydrolysis of cation reactions, especially for higher esters, where
methyl acetate,4-6 and the recovery of dilute acetic acid reaction kinetics is usually slower than that for the
from wastewater.7,8 The methyl acetate process not only methyl acetate system. Another advantage of the methyl
has been the first process employing the concept of acetate system is that the miscibility gap in the methyl
reactive distillation9 when esters gained commercial acetate-water system disappears upon the addition of
importance in the late 1920s10 but also was the first small amounts of methanol and/or acetic acid, which
process being adapted to a large industrial scale.2 Today, results in the absence of a second liquid phase in the
the hydrolysis of methyl acetate to acetic acid and reactive distillation column under normal operating
methanol is of major importance because methyl acetate conditions. This simplifies column modeling and allows
is a byproduct during the synthesis of poly(vinyl alcohol) the investigation of the backward reaction kinetics
under homogeneous conditions.
* To whom correspondence should be addressed. Phone: Because in modern reactive distillation processes a
+49-441-7983831.Fax: +49-441-7983330.E-mail: gmehling@tech.
structured or random packing is employed, it is desir-
chem.uni-oldenburg.de.
† Present address: Sulzer Chemtech Ltd., Separation and able to use a rate-based model. To apply a rate-based
Reaction Technology, P.O. Box 65, CH-8404 Winterthur, model for the simulation and design of a reactive
Switzerland. distillation process, a reliable knowledge of the reaction
10.1021/ie000063q CCC: $19.00 © 2000 American Chemical Society
Published on Web 06/17/2000
2602 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

kinetics and the chemical equilibrium is required be- Table 1. Size Distribution of Commercial Amberlyst 15
sides the mass-transfer characteristics, vapor-liquid diameter range (mm) mass fraction
equilibrium, and the various pure component properties.
>1.12 0.037
When the equilibrium-stage model is used, knowledge 1.00...1.12 0.211
of reaction kinetics is desirable to relax the often-made 0.80...1.00 0.403
simplifying assumption of chemical equilibrium. 0.63...0.80 0.242
The kinetics of this esterification reaction has been 0.40...0.63 0.102
investigated to some extent in the past. The reaction 0.20...0.40 0.004
<0.20 0.0003
has been studied in terms of heterogeneous catalysis
by various ion-exchange resins7,8,11-13 or homogeneous
catalysis by strong acids, like hydrogen chloride14 or 1). The calculated mean diameter of the resin beads was
hydrogen iodide.15 To the best of our knowledge, all 0.86 mm. For the investigation of possible mass-transfer
investigations published either covered only a limited resistance effects, a sample of the catalyst was ground
temperature range,13 covered a limited concentration and sieved to a diameter smaller than 63 µm.
range,8,14 or suffered from side reactions with the Measurements for the Determination of Chemi-
catalyst.15 In some of the published models, the back- cal Equilibrium. The experiments were carried out by
ward reaction has not been considered, which is a valid placing the desired amount of reactants (total 2.5 g,
assumption if one operates with a large excess of one different compositions) and vacuum-dry catalyst (about
reactant7,8,14 and at low conversion, but not when one 10 mg) in a 5-cm3 glass vial, which was then sealed and
operates at nearly stoichiometric concentrations, as in placed in a thermostat at a given temperature with an
the case of reactive distillation processes for the syn- accuracy of (0.1 K for 1-3 weeks. To check if the
thesis or the hydrolysis of methyl acetate. Our group chemical equilibrium was reached within analytical
already reported kinetic data for this system in the error, each experiment contained samples to undergo
past,3,16,17 which were fitted to a considerably smaller net esterification and samples to undergo net hydrolysis.
database than that in the present work. In this work, Some of these samples were composed in a way to give
we extend the range of the data from the stoichiometric the same equilibrium composition. Values reported
regime into the more diluted regions, as they are satisfy this criterion. Just before analysis, the vials were
encountered in reactive distillation processes, where the cooled rapidly to about 255 K and opened, and the liquid
reactants are fed in a countercurrent fashion into the phase was transferred into a 1-cm3 glass vial. This was
column. Also, a comparison of different kinetic models placed into the cooled (274 K) sample tray of an
and an independent investigation of chemical equilib- automatic sampler (HP 7673) and analyzed by gas
rium is presented. This work should therefore ensure a chromatography (HP 5890 with TCD; H2 as the carrier
reliable modeling of reaction kinetics in the methyl gas at 70 cm3‚min-1; Porapak-QS 1.25 m × 1/8 in.;
acetate system catalyzed by Amberlyst 15 or similar ion- temperature program, 383 K for 2.5 min, rate of 10
exchange resins. K‚min-1 to 443 K, held for 6.5 min; typical retention
times: water, 2.3 min, methanol, 4.5 min, methyl
Experimental Section acetate, 10.5 min, and acetic acid, 11.7 min).
Measurements for the Determination of Reac-
Chemicals. Acetic acid and methanol used were of tion Kinetics. Apparatus. The experiments were con-
analytical grade (99.8%, Scharlau), the water was ducted in a thermostated glass reactor with a volume
bidistilled, and the methyl acetate was reaction grade of 500 cm3. The temperature of the heating jacket was
(99.7% purity, as determined by gas chromatography, kept constant within (0.1 K. The temperature of the
Riedel-de-Haën). For the titrimetric analysis, Titrisol reaction mixture was measured by a thermocouple
(1.0 and 0.1 N sodium hydroxide, E. Merck) was used. against ice water. The stirrer was plate-type and the
The chemicals were used without further purification speed was variable between 100 and 800 rpm. To
except for drying over a molecular sieve. improve mixing, a baffle was installed. Furthermore, a
Catalyst. The macroreticular ion-exchange resin reflux condenser was used to avoid loss of volatile
Amberlyst 15 (wet) (Rohm and Haas Co.) was selected compounds. The outlets of the reactor not connected to
as the catalyst. The catalyst was washed once with the condenser were heated electrically to avoid conden-
methanol and then several times with water until the sation. Liquid samples of about 1 cm3 were taken using
supernatant liquid was colorless to remove impurities. a syringe and weighed (accuracy of the balance: (0.001
In some experiments, the catalyst was used wet, while g).
in others it was used after being dried. In all cases, the Experimental Procedure. Methanol, or in the case of
water content of the catalyst was determined after the the hydrolysis reaction water, was heated together with
kinetic experiment by drying of the catalyst at 90 °C the catalyst to the desired reaction temperature. Acetic
under vacuum until the mass remained constant (usu- acid, or for the reverse reaction methyl acetate, was
ally 2 days). Drying at higher temperatures involves the heated in a separate vessel. When the liquid in the
risk of loosing sulfonic acid sites in the form of SO3 reactor reached the desired reaction temperature, the
because of desulfonization of the polystyrene matrix of second reactant was added to the reactor. The stirrer
the catalyst. and the time measurement were started immediately.
The ion-exchange capacity of the resin was deter- Samples were then taken according to the sampling
mined by total exchange with sodium chloride solution program. During each experiment, between 15 and 45
and subsequent titration. A value of 4.77 ( 0.01 mequiv samples were taken.
of H+‚g-1 of dry catalyst was obtained, which compares Analysis. All reactants added to the reactor were
well with the value of 4.75 mequiv‚g-1 given by the weighed with an accuracy of (0.001 g. During the
manufacturer. A sieve analysis of the commercial resin experiments the acetic acid concentration of the samples
using mesh widths of 1.20, 1.12, 0.80, 0.63, 0.40, and was determined by potentiometric titration with sodium
0.20 mm gave the distribution of pellet diameters (Table hydroxide solution (Kyoto Electronics, model AT-200).
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2603

Table 2. UNIQUAC ri and qi Values19 Table 3. UNIQUAC Parameters Used for the Calculation
of Activity Coefficients (Eq 1)
component ri qi
i j aij (K) bij cij (K-1)
acetic acid 2.2024 2.0720
methanol 1.4311 1.4320 acetic acid methanol 390.26 0.970 39 -3.0613 × 10-3
methyl acetate 2.8042 2.5760 methanol acetic acid 65.245 -2.034 6 3.1570 × 10-3
water 0.9200 1.4000 acetic acid methyl -62.186 -0.436 37 2.7235 × 10-4
acetate
methyl acetic acid 81.848 1.116 2 -1.3309 × 10-3
With knowledge of the stoichiometry, the concentrations acetate
of the other reactants were calculated from the acetic acetic acid water 422.38 -0.051 007 -2.4019 × 10-4
acid concentration. The results were in good agreement water acetic acid -98.120 -0.29355 -7.6741 × 10-5
with preliminary gas-chromatographic determinations, methanol methyl 62.972 -0.710 11 1.1670 × 10-3
acetate
which also ensured the absence of side reactions. methyl methanol 326.20 0.724 76 -2.3547 × 10-3
Measurement Program. Fourteen runs for the reaction acetate
without catalyst and 50 runs for the reaction catalyzed methanol water -575.68 3.145 3 -6.0713 × 10-3
water methanol 219.04 -2.058 5 7.0149 × 10-3
by Amberlyst 15 were performed with variations in the methyl water 593.70 0.010 143 -2.1609 × 10-3
ratio of the reactants and the temperature (303.15- acetate
343.15 K). For the catalyzed runs, additionally, the water methyl acetate -265.83 0.962 95 2.0113 × 10-4
amount of catalyst was varied. The absence of external
mass-transfer resistance was ensured by a stirrer speed
sufficient to allow a complete suspension of the catalyst.
Additionally, the stirrer speed was varied between 100
and 560 rpm. No influence of the stirrer speed on the
kinetics could be detected above 170 rpm. Therefore, all
experiments reported were conducted with a stirrer
speed of at least 170 rpm, usually 250 rpm. The
influence of pore diffusion was checked in supplemen-
tary tests with sieved fractions of the catalyst having a
narrow distributed pellet diameter. Most of the experi-
ments were continued until chemical equilibrium was
reached.

Calculation of Liquid-Phase Activities


The activity coefficients, γi, necessary to account for
the real behavior of the liquid phase were calculated Figure 1. Chemical equilibrium constants Ka (O) and best fit of
using the UNIQUAC equation.18 A polynomial temper- our data (solid line), compared with an estimation based on eq 3
ature dependence of the interaction parameter and thermodynamic data (dashed line, short dashes). The correla-
tion given by Song et al.13 is included (dashed line, long dashes)
together with their mean values of Ka (0, from kinetic data) for
∆uij(T) ) aij + bijT + cijT2 (1) the individual temperatures. Song et al. included a value of Ka at
373 K given by Menschutkin28 in their correlation.
was included in the model. The r and q values of the
pure components (Table 2) were taken from the litera- Table 4. Standard Enthalpies and Standard Gibbs
Energies of Formation
ture.19 Interaction parameters were fitted simulta-
neously to vapor-liquid equilibrium (VLE), activity component source ∆h°f (kJ‚mol-1) ∆g°f (kJ‚mol-1)
coefficient at infinite dilution (γ∞), and heat of mixing acetic acid ref 20 -484.09 -389.23
(hE) data. The data originated from the Dortmund Data methanol ref 21 -239.11 -166.89
Bank, Version 1998, which was kindly placed at our methyl acetate ref 21 -442.79 -328.39
disposal by DDBST GmbH, Oldenburg, Germany. For water ref 20 -285.83 -237.14
the reactive systems (methanol-acetic acid and methyl
acetate-water), additional VLE and hE measurements and the standard Gibbs energy of reaction ∆g°r by
were carried out by our group. These data will be
reported elsewhere soon. The resulting parameters are ln Ka(T0) ) -∆g°r/RT0 (3a)
reported in Table 3.
and the integrated form of the van’t Hoff equation,

( )
Chemical Equilibrium Composition ∆h°r 1 1
To obtain reliable knowledge about the chemical ln Ka(T) ) ln Ka(T0) - - (3b)
R T T0
equilibrium, independent experiments were carried out
as described in the Experimental Section. The tabulated if a constant enthalpy of reaction is assumed, which is
results are available as Supporting Information. Figure reasonable because of the relatively small temperature
1 shows the results as a plot of the natural logarithm interval investigated.
of the equilibrium constant Ka This allows one to compare the results against an
estimation based on tabulated values of the standard
Ka ) ∏aiν ) ∏xiν × ∏γiν
i i i
(2) enthalpy of formation and the standard Gibbs energy
of formation of the individual reactants from the litera-
versus the inverse of the absolute temperature. The ture,20,21 which are given in Table 4. The resulting
slope and intercept of the linear regression of the data values for ∆h°r and ∆g°r are listed in Table 5. The
are related to the standard enthalpy of reaction ∆h°r dotted line in Figure 1 shows the resulting temperature
2604 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Table 5. Standard Enthalpies and Standard Gibbs Table 6. Summary of the Experiments without Catalyst
Energies of Reaction Added
source ∆h°r (kJ‚mol-1) ∆g°r (kJ‚mol-1) initial mole numbers
run temperature duration
from heats of formation (Table 4) -5.42 -9.41 number (K) (days) HOAc MeOH MeOAc H 2O
from linear regression (Figure 1) -5.67 -9.06 1 313.15 34 1.1840 1.1645 0 0
from Song’s correlation -6.51 -8.60 2 313.15 11 1.2934 0.6701 0 0
3 323.15 8 1.2150 1.2085 0 0
dependence of the equilibrium constant. However, one 4 323.15 8 1.4496 1.3914 0 0
should be aware that a small error in ∆g°f or ∆h°f 5 323.15 11 0.3814 0 1.7003 1.7597
causes a relatively large error in ∆g°r or ∆h°r due to 6 323.15 13 0.8944 0.2488 0.4011 0.5311
7 323.15 25 0.7933 0 0.5474 1.3108
the small magnitude of the values of these thermody- 8 328.15 20 2.7327 2.7278 0 0
namic properties. 9 333.15 3 1.4035 1.4033 0 0
The dashed line shows the correlation given by Song 10 333.15 6 1.2533 2.3720 0 0
et al.,13 along with their mean values of Ka for the 11 333.15 3 1.2102 0.5637 0.0710 0.1948
individual temperatures. The slopes of the three lines 12 333.15 15 0.8330 0 0.5641 1.2869
13 321.65 6 1.4373 0.1573 1.1350 1.3968
in Figure 1 are rather similar, indicating an agreement 14 321.65 1 1.4167 0.1317 1.1644 1.4417
of the different standard enthalpies of reaction. The
values are given in Table 5 together with the ∆g°r where xi denotes the mole fraction of component i, νi its
values, which also show good agreement.
stoichiometric coefficient, ai the activity of component i
(with ai ) xiγi), and k1 and k-1 the rate constants of the
Development of a Kinetic Model esterification and the hydrolysis reaction, respectively.
As a first step, the reaction kinetics without added The exponent R will be explained below in more detail.
catalyst is being modeled. The resulting model is then The temperature dependence of the rate constants is
incorporated into the model for the ion-exchange cata- expressed by Arrhenius’ law
lyzed reaction, to account for the homogeneously cata-

( )
lyzed reaction in the liquid phase. -EA,i
ki ) k0i exp (7)
Fitting Procedure
RT

The kinetic equations are integrated numerically by Four adjustable parameters, namely, the pre-exponen-
a fourth-order Runge-Kutta method. The calculated tial factors k01 and k-1
0
and the energies of activation
weight fractions of acetic acid are compared with the
EA,1 and EA,-1, have to be fitted. The term aRHOAc
experimental ones, giving the mean relative deviation
accounts for the activity of the protons in the mixture,
all samples

∑ | wHOAc,calc - wHOAc,exp
wHOAc,exp
| which are the catalyst for both the esterification and
the hydrolysis reaction. Depending on the assumptions
about the catalytic mechanism, R may show a value of
Frel ) × 100% (4) 0.5 or unity. On one hand, if one assumes that the
nsamples catalysis occurs via solvated protons and the dissociation
of acetic acid can be described by the Brønsted equilib-
and the mean squared deviation: rium constant Ka, then
all samples

∑ (wHOAc,calc - wHOAc,exp)2
Ka )
aH+aAcO- xH+xAcO-
≈ )
xH+2
(8)
Fabs2 ) (5) aHOAc aHOAc aHOAc
nsamples
and hence
Parameters are fitted by minimization of Fabs2 using the
Simplex-Nelder-Mead method,22 while the values of
Frel, obtained with the fitted parameters, are reported xH+ ≈ xaHOAcKa (9)
as residual error in Tables 7 and 11.
This assumption would lead to an exponent of R ) 0.5
Homogeneous Reaction in eq 6, when xKa is incorporated into the kinetic
In a first step, the homogeneous system was investi- constants.
gated. In this case, the reaction occurs by catalysis of On the other hand, if one assumes catalysis by
acetic acid. The esterification reaction and the hydroly- undissociated acetic acid, giving a reaction of two
sis reaction were investigated at temperatures between molecules of acetic acid with one molecule of methanol
313.15 and 333.15 K. A summary of the experiments is for the esterification reaction, and a reaction of one
given in Table 6. With the last two experiments, the molecule of methyl acetate, water, and acetic acid for
chemical equilibrium composition was verified because the hydrolysis reaction, this would lead to an exponent
even after weeks, chemical equilibrium was not reached. of R ) 1 in eq 6. Formally, this corresponds to a
The resulting composition profiles versus time were trimolecular reaction.
used to fit the parameters of the following kinetic To distinguish between these two possibilities, all
equation, data were fitted once with a value of R ) 0.5 and once
with R ) 1. Finally, R was included in the fitting
1 dxi
r) ) aRHOAc(k1aHOAcaMeOH - k-1aMeOAcaH2O) (6) procedure as a fifth adjustable parameter. All optimiza-
νi dt tions have been run with the assumption of ideal liquid-
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2605

Table 7. Parameters and Residual Errors of the Different Kinetic Models Used To Fit the Experimental Data without
Catalyst Added
mean relative
exponent R activities error (%) k01 (s-1) EA,1 (kJ‚mol-1) 0
k-1 (s-1) EA,-1 (kJ‚mol-1)
0.5 ideal 4.9 6.57 × 104 60.4 1.02 × 107 79.1
0.5 UNIQUAC 4.0 1.30 × 105 61.6 1.00 × 107 82.2
1 ideal 3.1 2.54 × 106 62.5 1.00 × 107 77.3
1 UNIQUAC 2.2 5.11 × 105 63.5 9.83 × 106 80.2
1.406 ideal 2.2 2.42 × 106 67.3 9.27 × 106 75.8
1.056 UNIQUAC 2.2 6.06 × 105 63.8 9.84 × 106 80.0

Figure 2. Arrhenius diagram of the rate constants for the


esterification reaction k1 (b) and the hydrolysis reaction (O) of the Figure 3. Experimental values for the acetic acid concentration
homogeneous reaction catalyzed by acetic acid (eq 6, R ) 1, activity (weight fraction) of four runs without catalyst added (0, run 3; 4,
coefficients calculated by UNIQUAC). The lines represent the run 5; ), run 6; O, run 7; compare Table 6) and calculated course
result of the overall fit (compare Table 7). (eq 6, R ) 1, activity coefficients calculated by UNIQUAC).

phase behavior (γi ) 1.0, this means ai ) xi) and with UNIQUAC model. This emphasizes the importance of
values for γi calculated from the UNIQUAC equation taking into account the real liquid-phase behavior to
and thus accounting for the real liquid-phase behavior. obtain consistent results. For consistency, we recom-
The results of the fitting procedure are reported in Table mend using UNIQUAC and a value of R ) 1. This
7. To ensure that the global minimum was found, the corresponds formally to catalysis by molecular acetic
optimization was restarted at least three times with acid, not solvated protons. This result is consistent with
different initial parameters. results published by Williamson and Hinshelwood.23
As can be seen from Table 7, the best results for a
fixed value of R were obtained with a value of R ) 1 for Heterogeneously Catalyzed Reaction
ideal and nonideal behavior. Accounting for the real
liquid-phase behavior always resulted in a smaller Preliminary experiments were used to determine a
residual error. For the case of R ) 1 and liquid-phase sufficient stirrer speed to ensure the absence of external
activities being calculated by UNIQUAC, the validity mass-transfer limitations. Additionally, some experi-
of the kinetic model has been assessed by fitting only ments to assess intraparticle mass transfer were con-
k1 and k-1 to all experimental data at the same ducted by using different sieve fractions of the commer-
temperature. Without experiments 13 and 14 from cial catalyst and using sieved samples of ground catalyst.
Table 6, for which no meaningful fit was obtained For sieve fractions of the commercial catalyst, no effect
because of the near-absence of net reaction, four pairs of the pellet size on the reaction kinetics was observed,
of k1/k-1 values for the different temperatures were which is consistent with the fact that pellets of Am-
obtained. The natural logarithm of these values is berlyst 15 are composed of very small microspheres,24
plotted in Figure 2 versus the inverse temperature, thus which are similar in size. When the catalyst was ground
giving an Arrhenius plot. Rate constants calculated from in a ceramic mill and sieved to smaller than 63 µm, an
the overall fit (row 4 in Table 7) are included as lines. increase in the rate by a factor of about 2 was observed.
From the plot, it is obvious that the temperature This indicates indeed a mass-transfer limitation, which
dependence of the rate constants can be described by is only of importance when the catalyst is subjected to
Arrhenius’ law and that the overall fit is able to grinding or abrasion. To avoid abrasion, no magnetic
represent the data. To demonstrate that the model is stirrer, but a plate-type stirrer, was used in our work.
capable of giving a reasonable representation of the Neither the data of Song et al.,13 which were obtained
experimental data, four experiments at 323.15 K with using a magnetic stirrer, nor our data obtained with a
very different initial compositions are plotted in Figure magnetic stirrer or with a ground catalyst were included
3 together with the calculated concentration course. As in our fitting procedure. The error introduced by the
can be seen from the figure, the model is able to describe abrasion of the resin should be small, though, because
not only the esterification but also the hydrolysis the mass fraction of the smaller particles is relatively
reaction. low. Because the aim of this work is to present a kinetic
When R was fitted, a value of about unity was expression for use in reactive distillation, where the
obtained, taking into account the real behavior by the catalyst is packed into structured packing elements and
2606 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Table 8. Summary of the Experiments with Amberlyst 15 as Catalyst

run mass of dry initial mole numbers (mol)


number temperature (K) catalyst (g) duration (h) HOAc MeOH MeOAc H2O
1 304.25 18.27 23.4 1.557 1.557 0 1.014
2 303.45 31.54 23.3 1.993 2.012 0 0
3 313.63 8.84 2.0 1.621 1.618 0 0.542
4 313.14 3.81 23.4 1.646 1.608 0 0.234
5 313.00 5.76 31.6 3.589 1.591 0 0.354
6 323.00 7.17 5.6 1.997 1.999 0 0.427
7 323.15 7.71 6.4 2.031 1.982 0 0
8 323.00 7.31 4.8 1.996 2.090 0 0.441
9 323.00 6.70 5.0 2.006 2.017 0 0.363
10 323.00 23.80 5.1 2.003 2.049 0 1.485
11 323.08 8.18 17.7 1.656 1.626 0 0.502
12 323.06 8.55 18.8 1.642 1.621 0 0.524
13 323.06 17.10 21.1 1.632 1.627 0 1.049
14 323.00 3.54 23.0 2.050 1.019 0 0.217
15 323.00 8.67 3.2 1.539 2.921 0 0.531
16 323.15 7.15 5.9 0.428 3.588 0 0.440
17 333.14 4.25 23.2 1.669 1.642 0 0.261
18 333.12 8.19 20.3 1.670 1.646 0 0.502
19 333.13 17.71 20.1 1.661 1.604 0 1.087
20 333.20 23.38 4.9 1.562 1.549 0 1.230
21 304.25 18.66 23.4 0 0.367 1.614 1.546
22 313.30 17.61 19.7 0 0 1.621 2.801
23 313.00 5.70 95.1 0 0 2.539 4.156
24 323.08 17.13 18.5 0 0 1.648 2.698
25 323.15 29.52 3.7 0 0.432 1.812 1.897
26 333.03 17.78 20.9 0 0 1.623 2.785
27 323.15 12.97 6.5 2.001 2.009 0 0
28 323.15 27.61 6.5 2.012 2.008 0 0
29 323.15 36.95 6.5 2.003 1.992 0 0
30 323.15 46.35 6.5 2.015 1.998 0 0
31 323.15 55.37 3.7 2.004 2.015 0 0
32 323.15 5.02 6.5 0.246 4.761 0 0
33 323.15 7.20 6.2 0.431 3.579 0 0
34 323.15 6.14 7.1 1.008 4.073 0 0
35 323.15 5.61 5.4 2.721 0.730 0 0
36 323.15 11.19 6.2 3.590 0.387 0 0
37 323.15 5.89 1.8 3.088 0.142 0 0
38 323.15 4.32 70.6 0 0 0.595 5.400
39 323.15 4.35 24.6 0 0 1.005 3.991
40 323.15 6.65 5.4 0 0 1.997 0.499
41 323.15 5.27 4.5 0 0 1.795 0.198
42 313.15 20.73 24.5 0 0 0.898 9.362
43 323.15 20.54 5.4 0.897 0.901 0 8.454
44 323.15 59.68 1.3 0 0 0.613 5.417
45 333.15 20.49 5.4 0 0 0.904 9.366
46 343.15 21.18 3.5 0.894 0.901 0 8.517
47 313.49 17.60 21.1 1.483 0 0.640 1.477
48 313.45 17.58 19.0 1.012 0 0.956 2.328
49 333.05 17.66 15.8 0.654 1.580 0.800 1.281
50 333.09 17.58 19.0 0.459 0.149 0.934 3.866
thus not subjected to abrasion, it is justified to neglect hydrolysis reaction (Table 8). Besides the temperature
this possible mass-transfer effect. The reader should be for the esterification (experiments 1-20) and hydrolysis
aware that all activation energies for the heteroge- (experiments 21-26) reaction, the amount of catalyst
neously catalyzed reaction presented in this work are (experiments 27-31), the initial reactant ratio for the
apparent activation energies, not the true values for the esterification (experiments 32-37), and the hydrolysis
chemical reaction. (experiments 38-41) reaction were varied at 323 K.
We also checked the applicability of the model pre- Finally, some experiments (42-46) were performed in
sented by Song et al.13 by calculating the concentration the presence of excess water, which permitted the
course of acetic acid from their model and comparing temperature to be raised to 343 K without boiling, and
them with our experimental results. Their model de- for initial compositions near chemical equilibrium (ex-
scribes the reaction quite well when the experimental periments 47-50).
conditions are similar to theirs, that is, esterification According to the data presented by Song et al.,13 the
reaction and in the presence of 8-22 wt % of wet side reaction, the condensation of two molecules of
catalyst. However, the hydrolysis reaction is not de- methanol to form dimethyl ether, does not occur in an
scribed by their model with the required accuracy, as appreciable amount in the temperature range covered
are experiments with considerably less catalyst or dry in this work and is therefore neglected.
catalyst. Therefore, it seems appropriate to propose a Thereactionhasbeendescribedbothwithapseudohomo-
model with parameters based on a larger experimental geneous model6-8,11 and with adsorption-based models.13
database without these limitations. Therefore, a comparison of both models is presented.
Our database includes experimental data from 303 On one hand, theoretical support for the pseudohomo-
to 333 K for both the esterification reaction and the geneous model comes from the fact that the polymeric
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2607

catalyst swells in contact with polar solvents such as Table 9. Experimental Swelling Ratios Obtained for the
water or methanol by more than 50% of its dry volume. Pure Components and Adsorbed Volumes, Masses, and
This should facilitate diffusion in the polymer matrix Moles per Gram of Dry Amberlyst 15 at 298 K
and make the polymer-bound sulfonic acid groups adsorbed adsorbed adsorbed
readily accessible for the reactants. On the other hand, swelling volume mass amount
the swelling is highly selective,13,25 which was confirmed component ratio (cm3‚g-1) (g‚g-1) (mmol‚g-1)
by our experiments. This makes an adsorption-based acetic acid 1.43 0.307 0.319 5.31
model suitable, even if on a molecular scale, adsorption methanol 1.55 0.393 0.309 9.60
seems not to be the best description for the phenomenon. methyl acetate 1.40 0.286 0.265 3.58
water 1.67 0.479 0.478 26.5

Pseudohomogeneous Model by the dry volume of the resin, in a sealed graduated


cylinder (accuracy, (0.25 cm3) are reported in Table 9.
First of all, the pseudohomogeneous model should be The values given are mean values of four to five
considered, which can be written as experiments conducted at 298 K. The accuracy of the
data is estimated to be about (5%. Assuming that the
1 dni resin bed’s volume changes in the same way as the
r) ) mcat(k1aHOAcaMeOH - k-1aMeOAcaH2O) (10) volumes of the individual resin microspheres, adsorbed
νi dt
volumes can be calculated from the swelling ratio. Also
assuming ideal mixing, that is, no excess volume, the
where the ion-exchange resin is thought to act as a adsorbed mass and amount per gram of catalyst can be
source of solvated protons. As for the homogeneous calculated. The values are given in Table 9.
reaction, the model parameters were fitted to the It is obvious from Table 9 that the assumption of a
experimental data by the Simplex-Nelder-Mead method, constant amount adsorbed is not suitable because there
minimizing Fabs2. The resulting parameters are given is an order of magnitude difference between the values
in Table 11 both for assuming ideal liquid-phase be- for methyl acetate and water. A better choice would be
havior and considering liquid-phase activities using the the assumption of constant adsorbed mass, for which
UNIQUAC model. Also, the results of calculating the the values for the individual compounds are similar. For
concentration course with the model of Song et al.13 are this reason, the equations given by Song et al. are
given. For this model, the calculation was performed modified to allow for a constant adsorbed mass.
with the parameters reported by Song et al. and also From the overall mass balance of the binary liquid-
with parameters fitted to our data. phase adsorption in a system of total solvent weight,
The pseudohomogeneous model is clearly not the best m0, and catalyst mass, mcat, with the overall weight
choice. An adsorption-based model, such as the model
fraction, w01, and the equilibrium liquid-phase weight
of Song et al., seems to be a better choice because the
residual error can be halved when compared with the fraction, wL1 , of solvent 1, it follows for the adsorbed
pseudohomogeneous model, when the parameters of the masses of solvents 1 and 2, mS1 and mS2 , analogous to
model of Song et al. are fitted to our data. Kipling27

Adsorption-Based Model
m0(w01 - wL1 ) mS1 wL2 - mS2 wL1
) (11)
mcat mcat
A rigorous modeling of the solvent uptake by a
swelling polymer would involve modeling of the polymer- Assuming Langmuir-type adsorption based on mass, the
phase activities by an appropriate model, for example, following relation for the mass coverage mSi /mS can be
a modified Flory-Huggins model.25 The condition of obtained, with mS being the total adsorbed mass and ai
isoactivity between the liquid and polymer phase would the liquid-phase activities:
then yield the swelling equilibrium. However, for densely
cross-linked resins bearing highly polar groups on
mSi Kiai
almost every monomer (Amberlyst 15 has a degree of
functionalization of about 88%), these effects are poorly
) (12)
understood and do not yield consistent results.25 For this mS 1+ ∑j Kjaj
reason, we will stay with the adsorption-based modeling
until a reliable model applicable to this system has been
developed. Combining eqs 11 and 12 for the binary case, the
To further improve the modeling of the reaction following relation is obtained,
kinetics, the Langmuir-Hinshelwood-Hougen-Watson
(LHHW) model of Song et al. should be modified. A m0(w01 - wL1 ) L
mS K1a1w2 - K2a2w1
L
) (13)
critical assumption in their adsorption model is that a mcat mcat 1 + K1a1 + K2a2
constant adsorbed amount is assumed, the value of
which is not given in the article. Furthermore, they did which can readily be used to fit the equilibrium con-
not use a constant value for the fit of all binary stants Ki and the total adsorbed mass mS/mcat to binary
adsorption data because they fitted this value for each adsorption data, for example, by the Simplex-Nelder-
binary system separately.26 It is necessary to fit the Mead method.22 This is done with the four nonreactive
adsorption constants to independent data.13 If they are sets of binary adsorption data, which were also obtained
fitted together with the kinetic constants, an extremely using a sealed graduated cylinder to obtain the swelling
good fitting result with very poor extrapolation capabili- ratio information, which is not included in this model
ties is obtained. but might be useful for developing a more reliable
The results of our measurements of the swelling ratio, polymer swelling model. The total adsorbed mass per
which is the volume of solvent-equilibrated resin divided catalyst mass, mS/mcat, is kept the same for all binary
2608 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

Figure 4. Relative adsorption of water from water-acetic acid Figure 6. Arrhenius diagram of the rate constants for the
(O) and water-methanol (0) mixtures on Amberlyst 15 and esterification reaction k1 (b) and the hydrolysis reaction (O) of the
calculated dependence assuming constant adsorbed mass. heterogeneously catalyzed reaction (eq 16, activity coefficients
calculated by UNIQUAC). The lines represent the results of the
overall fit (compare Table 11).

1 dni
r) ) mcat(k1xSHOAcxSMeOH - k-1xSMeOAcxH
S
2O
) (14)
νi dt

where

mSi /Mi
xSi ) (15)
∑j (mSj /Mj)

represents the mole fractions of the reactants in the


adsorbate phase yields

Figure 5. Relative adsorption of methanol from a mixture of


methanol and methyl acetate (O) and relative adsorption of acetic
acid from a mixture of acetic acid and methyl acetate (0) on
Amberlyst 15 and calculated dependence assuming constant
adsorbed mass.
r)
1 dni
νi dt
) mcat
(
k1a′HOAca′MeOH - k-1a′MeOAca′H2O
(a′HOAc + a′MeOH + a′MeOAc + a′H2O)2 )
Ki ai
Table 10. Results of the Regression for the Nonreactive with a′i ) (16)
Binary Adsorption Data at 298 K
Mi
Equilibrium Adsorbed Mass with Mi denoting the molar mass of component i,
mS/mcat ) 0.95
whereby the rate constants k1 and k-1 are temperature
Adsorption Equilibrium Constants dependent, as described by eq 7.
KHOAc ) 3.15 For this model, the Arrhenius diagram shown in
KMeOH ) 5.64
KMeOAc ) 4.15
Figure 6 has been constructed by fitting simultaneously
KH2O ) 5.24 all experimental data at a given temperature and
plotting the natural logarithm of the resulting values
for ki against the reciprocal temperature. Also included
systems. The data are presented in Figures 4 and 5 and in the plot is the result for the fit of the four variables
are available as Supporting Information. The results of k01, k-1
0
, EA,1, and EA,-1 as lines, the result of which is
a fit of all four adsorption equilibrium constants to all given in the last row of Table 11. The agreement of the
of the data are included in Figures 4 and 5 and overall fit and the results obtained for the individual
summarized in Table 10. The fit reproduces the data temperatures is good, as is the linearity.
within experimental accuracy. The comparatively high Figure 7 shows results of a fit of eq 16 to single
value obtained for mS/mcat might be explainable by the experiments, where only k1 was fitted and k-1 was
increase in the swelling ratio when the liquid phase is calculated from our chemical equilibrium correlation by
a mixture instead of a pure solvent. the relation

In contrast to the work of Song et al.,13 all four aMeOAcaH2O k1KHOAcKMeOHMMeOAcMH2O


adsorption constants are known at this point and none Ka ) ) (17)
aHOAcaMeOH k-1KMeOAcKH2OMHOAcMMeOH
of them needs to be included as adjustable parameters
in the fit to the kinetic data. Thus, the adsorption is
being modeled truly independent from the kinetic fit. which can be obtained from eq 16 when the reaction rate
r is set to zero.
Combining this adsorption model (eq 12) with a In these experiments, the mass of added catalyst was
kinetic model of the form varied to ensure the applicability of the model also at
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2609

Table 11. Parameters and Residual Errors of the Different Kinetic Models Used To Fit the Experimental Data with
Amberlyst 15 as the Catalyst
mean relative k01 EA,1 k-10 EA,-1
model error (%) (mol‚g-1‚s-1) (kJ‚mol-1) (mol‚ g-1‚s-1) (kJ‚mol-1)
pseudohomogeneous, 14.6 1.648 × 104 47.98 1.161 × 105 58.60
ideal liquid phase
pseudohomogeneous, 13.9 2.961 × 104 49.19 1.348 × 106 69.23
activities by UNIQUAC
Song’s model with original parameters, 10.4 3.985 × 106 52.28 use Ka given by Song et al.
activities by UNIQUAC
Song’s model, parameters fitted, 6.7 5.312 × 107 58.25 use Ka from this work
activities by UNIQUAC
eq 16 5.4 8.497 × 106 60.47 6.127 × 105 63.73

Figure 7. Rate constant of the esterification reaction k1 (b) at Figure 9. Experimental values (O) for the acetic acid concentra-
323 K versus catalyst mass for experiments 7 and 27-31. Only k1 tion and predicted course by the pseudohomogeneous model (eq
has been fitted for each run using the known chemical equilibrium 10, dashed lines, activity coefficients by UNIQUAC) and eq 16
constant from Figure 1. The line represents the result of the overall (solid lines). Five runs (experiments 25 and 38-41) of the
fit (compare Table 11). heterogeneously catalyzed hydrolysis reaction with varying initial
reactant ratios are shown.

Table 12. Standard Enthalpies and Standard Gibbs


Energies of Reaction Obtained from the Kinetic
Parameters
kinetic model ∆h°r (kJ‚mol-1) ∆g°r (kJ‚mol-1)
pseudohomogeneous, -20.04 -10.57
activities by UNIQUAC
eq 16 -3.26 -8.37

homogeneous model. Both models use the same number


of adjustable parameters for the fit, that is, k01, k-1
0
,
EA,1, and EA,-1 because the adsorption constants Ki have
been fitted independently to binary adsorption data. An
attempt of fitting the four kinetic parameters and the
four adsorption constants to kinetic data only resulted
in very small residual errors, but also in a very poor
extrapolability of the model to experimental data not
Figure 8. Experimental values (O) for the acetic acid concentra- included in the fitting procedure.
tion and predicted course by the pseudohomogeneous model (eq The consistency between kinetics and chemical equi-
10, dashed lines, activity coefficients by UNIQUAC) and eq 16
(solid lines). Seven runs (experiments 7 and 32-37) of the
librium can be checked by calculating Ka from the
heterogeneously catalyzed esterification with varying initial re- following equation,
actant ratios are shown.
aMeOAcaH2O k1
high catalyst concentrations. It can be seen from Figure Ka ) ) (18)
7 that the assumption of a linear dependence of the rate aHOAcaMeOH k-1
on the mass of catalyst is valid.
To demonstrate the validity of the model in the for the pseudohomogeneous model with activities and
diluted concentration range, the concentration courses from eq 17 for the adsorption-based model. If Ka is
of experiments with varying initial molar ratios are substituted by eq 3, the heats of formation can be
presented in Figure 8 for the esterification reaction and calculated from the kinetic parameters in Table 11. The
in Figure 9 for the hydrolysis reaction. Also included is results are shown in Table 12. A comparison with Song’s
the concentration course calculated from eq 16 and from model is not possible because it needs the adsorption
the pseudohomogeneous model. Obviously, eq 16 de- constant of methyl acetate as an additional parameter
scribes the kinetic data better than the simple pseudo- and, furthermore, the chemical equilibrium constant Ka
2610 Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000

is included in the kinetic equation. In a comparison with adsorption experiments (24 pages) is available free of
Table 5, one can see how obvious it is that the value for charge via the Internet at http://pubs.acs.org.
∆h°r obtained from the adsorption-based kinetics shows
better agreement between kinetics and thermodynamics Literature Cited
than that of the pseudohomogeneous model. The values
for ∆g°r agree well for both kinetic models, meaning (1) Bessling, B.; Loening, J.-M.; Ohligschlaeger A.; Schembecker
G.; Sundmacher, K. Investigations on the Synthesis of Methyl
that the magnitude of Ka at 298.15 K, which depends Acetate in a Heterogeneous Reactive Distillation Process. Chem.
only on ∆g°r, is represented well by both kinetic mod- Eng. Technol. 1998, 21, 393.
els, but the temperature dependence of Ka, which is (2) Agreda, V. H.; Partin, L. R.; Heise, W. H. High-Purity
governed by ∆h°r, deviates somewhat from the value Methyl Acetate via Reactive Distillation. Chem. Eng. Prog. 1990,
obtained from thermodynamic information. 86 (Feb), 40.
(3) Krafczyk, J.; Gmehling, J. Einsatz von Katalysatorpackun-
gen für die Herstellung von Methylacetat durch reaktive Rekti-
Conclusions fikation. Chem.-Ing.-Tech. 1994, 66, 1372.
(4) Fuchigami, Y. Hydrolysis of Methyl Acetate in Distillation
The design of reactive distillation columns requires Column Packed with Reactive Packing of Ion Exchange Resin.
reliable kinetic data besides other information, like Chem. Eng. Jpn. 1990, 23, 354.
phase equilibrium data, chemical equilibrium data, (5) Han, S. J.; Jin, Y.; Yu, Z. Q. Application of a Fluidized
Reaction-Distillation Column for Hydrolysis of Methyl Acetate.
transport property data, or the various pure component Chem. Eng. J. 1997, 66, 227.
properties. In this work, kinetic expressions for the (6) Ge, X.; Wang, J.; Wang, Z.; Jin, Y. Ein allgemeingueltiges
methyl acetate system are presented with parameters Modell zur Simulation der katalytischen Destillation am Beispiel
fitted to a fairly large database covering both the der Hydrolyse von Methylacetat. Chem.-Tech. (Leipzig) 1999, 51,
esterification and the hydrolysis reaction. For the 69.
homogeneous reaction, the kinetic fit is improved by (7) Neumann, R.; Sasson, Y. Recovery of Dilute Acetic Acid by
Esterification in a Packed Chemorectification Column. Ind. Eng.
using activities instead of mole fractions. The reaction
Chem. Process Des. Dev. 1984, 23, 654.
follows a rate expression that suggests catalysis by (8) Xu, Z. P.; Chuang, K. T. Kinetics of Acetic Acid Esterification
molecular acetic acid instead of by solvated protons. This over Ion Exchange Catalysts. Can. J. Chem. Eng. 1996, 74, 493.
is consistent with the results from Williamson and (9) Backhaus, A. A. Continuous Process for the Manufacture
Hinshelwood.23 The magnitude of the homogeneous of Esters. U.S. Patent 1,400,849, 1921.
reaction rate shows that this reaction can be neglected (10) Keyes, D. B. Esterification Processes and Equipment. Ind.
in reactive distillation columns, unless very high liquid Eng. Chem. 1932, 24, 1096.
(11) Nageshwar, G. D.; Pandharpurkar, D. M.; Mene, P. S.
holdups are present. Esterification of Methanol with Acetic Acid-Catalyzed by a Cation
For the heterogeneously catalyzed reaction, the use Exchange Resin. Indian J. Technol. 1975, 14, 310.
of activities alone in a pseudohomogeneous model (12) Zidan, F.; El-Hadi, M. F.; El-Nahas, M. R. Ion Exchange
improves the fit only slightly. Taking the liquid-phase Resin-Catalysed Esterification of Aliphatic and Aromatic Acids
nonideality into account is obviously not sufficient to with Methanol and Ethanol. Pakistan J. Sci. Res. 1986, 29, 397.
(13) Song, W.; Venimadhavan, G.; Manning, J. M.; Malone, M.
describe the heterogeneous kinetics. The interaction
F.; Doherty, M. F. Measurement of Residue Curve Maps and
between the solid catalyst and the reactants has to be Heterogeneous Kinetics in Methyl Acetate Synthesis. Ind. Eng.
considered in the model. Despite the fact that a precise Chem. Res. 1998, 37, 1917.
and thermodynamically consistent treatment of systems (14) Smith, H. A. Kinetics of the Catalyzed Esterification of
involving swelling ion-exchange resins is not available, Normal Aliphatic Acids in Methyl Alcohol. J. Am. Chem. Soc. 1939,
a good fit of the kinetic experiments can be obtained 61, 254.
using an adsorption-based model. Following the concept (15) Roennback, R.; Salmi, T.; Vuori, A.; Haario, H.; Lehtonen,
J.; Sundqvist, A.; Tirronen, E. Development of a Kinetic Model
of Song et al.,13 independent adsorption experiments for the Esterification of Acetic Acid with Methanol in the Presence
were used to determine the adsorption equilibrium of a Homogeneous Acid Catalyst. Chem. Eng. Sci. 1997, 52, 3369.
constants which are then incorporated into the kinetic (16) Götze L. Makrokinetik der heterogen katalysierten Ver-
expression. Including adsorption information leads to esterung von Methanol und Essigsäure sowie der Hydrolyse von
a much better description of the kinetic results. Using Methylacetat in strukturierten Packungselementen. Ph.D. Dis-
an adsorption expression based on adsorbed mass sertation, Carl-von-Ossietzky University, Oldenburg, Germany,
1998.
instead of adsorbed amount results in further improve- (17) Pöpken, T.; Geisler, R.; Götze, L.; Brehm, A.; Moritz, P.;
ment of the fitting result. The resulting model (eq 16) Gmehling, J. Reaction Kinetics and Reactive DistillationsOn the
can easily be incorporated into commercial simulators. Transfer of Kinetic Data from a Batch Reactor to a Trickle-Bed
Additionally, this model provides better agreement Reactor. Chem. Eng. Technol. 1999, 21, 401.
between kinetics and thermodynamics (chemical equi- (18) Abrams, D. S.; Prausnitz, J. M. Statistical Thermodynam-
librium) than the pseudohomogeneous model. ics of Liquid Mixtures: A New Expression for the Excess Gibbs
Energy of Partly or Completely Miscible Systems. AIChE J. 1975,
21, 116.
Acknowledgment (19) Gmehling, J.; Onken, U. Vapor-Liquid Equilibrium Data
Collection. In DECHEMA Chemistry Data Series; DECHEMA:
We are grateful for the generous financial support Frankfurt, Germany, 1977; Vol. 1, Part 1.
provided by Sulzer Chemtech Ltd. (Winterthur, Swit- (20) Barin, I. Thermochemical Data of Pure Substances; VCH-
Verlagsgesellschaft: Weinheim, Germany, 1989.
zerland) and the “Fonds der Chemischen Industrie” by (21) Domalski, E. S.; Hearing, E. D. Estimation of the Ther-
granting a scholarship to T.P. We also would like to modynamic Properties of C-H-N-O-S-Halogen Compounds at
thank Robert Geisler, Peter Kunze, and Sven Steini- 298.15 K. J. Phys. Chem. Ref. Data 1993, 22, 805.
geweg for performing some of the kinetic experiments. (22) Nelder, J. A.; Mead, R. A Simplex Method for Function
Minimization. Comput. J. 1967, 7, 308-313.
(23) Williamson, A. T.; Hinshelwood, C. N. The Kinetics of
Supporting Information Available: Tables of Esterification. The Reaction between Acetic Acid and Methyl
experimental data for the chemical equilibrium compo- Alcohol Catalysed by Hydrions. Trans. Faraday Soc. 1934, 30,
sition, the different kinetic runs, and the results of the 1145.
Ind. Eng. Chem. Res., Vol. 39, No. 7, 2000 2611

(24) Pitochelli, A. R. Ion Exchange Catalysis and Matrix Effects; (28) Menschutkin, N. Ueber den Einfluss der Isomerie der
Rohm and Haas Co.: Philadelphia, PA, 1980. Alkohole und der Saeuren auf die Bildung zusammengesetzter
(25) Mazzotti, M.; Neri, B.; Gelosa, D.; Krugov, A.; Morbidelli, Aether. Liebigs Ann. Chem. 1879, 195, 334-364.
M. Kinetics of Liquid-Phase Esterification Catalyzed by Acidic
Resins. Ind. Eng. Chem. Res. 1997, 36, 3-10. Received for review January 19, 2000
(26) Doherty, M. F. University of Massachusetts at Amherst, Revised manuscript received April 6, 2000
personal communication, 2000. Accepted April 11, 2000
(27) Kipling J. J. Adsorption From Solutions of Non-Electro-
lytes; Academic Press: London, 1965. IE000063Q

You might also like