Você está na página 1de 6

4060

Ind. Eng. Chem. Res. 2004, 43, 4060-4065

KINETICS, CATALYSIS, AND REACTION ENGINEERING


Thermodynamics and Kinetics of the Dehydration of tert-Butyl
Alcohol
Maija L. Honkela,* Tuomas Ouni, and A. Outi I. Krause
Helsinki University of Technology, Department of Chemical Technology, P.O. Box 6100,
Espoo FIN-02015 HUT, Finland

The dehydration of tert-butyl alcohol to water and isobutene was studied using an ion-exchange
resin catalyst at temperatures between 60 and 90 C. A temperature-dependent equilibrium
constant for the dehydration reaction was obtained that gave a reaction enthalpy of 26 kJ mol-1,
in good agreement with values in the literature. Measured data were used for kinetic modeling
of the reaction. The best model with physically meaningful parameters was of LangmuirHinshelwood type where isobutene does not adsorb on the catalyst. The activation energy for
the reaction in this case was 18 kJ mol-1.
Introduction
Both the dehydration of tert-butyl alcohol (TBA) to
isobutene and the reverse reaction have been studied
widely. TBA production is of interest because of the use
of TBA as a gasoline component [RON (research octane
number) ) 109, MON (motor octane number) ) 91].1
TBA formed as a side product in 1,2-epoxypropane
synthesis is used not only as a gasoline component2 but
also in the production of isobutene for methyl tert-butyl
ether (MTBE) and other high-octane gasoline components.3,4 Also, direct routes from TBA to ethers (without
isobutene formation in between) exist.5-7
Gates et al.8 studied TBA dehydration on Amberlyst
15 ion-exchange resin catalyst and proposed a carbonium ion mechanism at low TBA concentrations. In
this mechanism TBA and the proton of the catalyst form
a tert-butyl cation. This cation can either react back to
TBA or form isobutene at the same time as the proton
is regenerated. At higher concentrations, the reaction
was reasoned to proceed according to a concerted
mechanism involving the participation of several active
sites. In later studies, Gates and Rodriguez9 proposed
a rate equation that takes into account the active sites
that result from alcohol adsorption. In the presence of
water, the reaction was of first order with respect to
the TBA concentration, and with low water contents,
the rates were represented by Langmuir-Hinshelwoodtype kinetics. Abella et al.10 also studied the kinetics of
the dehydration of TBA in the liquid phase. However,
they used an atmospheric-pressure system, and because
the isobutene that was produced evaporated from the
mixture, the reaction was considered irreversible.
The hydration of isobutene on ion-exchange resins is
different from the dehydration of TBA because a large
amount of water is present on the catalyst. Gupta and
Douglas,11 for example, carried out experiments in
which water was present in large excess so that the
* To whom correspondence should be addressed. E-mail:
maija.honkela@hut.fi. Fax: +358-9-451 2622.

resin was fully swollen. They obtained first-order irreversible kinetics for the hydration reaction.
Delion et al.12 applied various solvents in the hydration of isobutene with the aim of keeping the mixture
in a single liquid phase. They tested p-dioxane, acetone,
nitromethane, butylcellosolve (2-butoxyethanol), isopropyl alcohol, cyclohexanol, tetrahydrofurfurylic alcohol,
and acetic acid and calculated solvent-dependent equilibrium constants for the reaction. Velo et al.13 obtained
both equilibrium constants and kinetics for the hydration of isobutene. The kinetic equations were based on
a carbonium ion mechanism in which isobutene forms
a tert-butyl cation with the proton of the catalyst. They
also concluded that TBA inhibits the hydration of
isobutene more than water. In another study,14 they
found that intraparticle diffusivity increased with temperature and decreased with TBA concentration.
Diffusion has also been studied in other publications.
TBA dehydration studies indicate that, when macroporous ion-exchange resins are used as catalysts, masstransport limitations do not exist.8,15 Mass transport
seems to affect the rates of isobutene hydration, however.16 Studies in a trickle-bed reactor with an aqueous
phase and isobutene as the gas indicated that intraparticle diffusion has a greater effect on the rate than
does liquid-to-particle mass transfer.17 In trickle-bed
reactors, both the wetting efficiency and mass transfer
influence the total rate.18
On an acidic ion-exchange resin catalyst, isobutene
reacts to form diisobutenes and higher oligomers.
Moreover, when diisobutenes need to be produced
selectively, polar components such as TBA, water, or
methanol are added to the dimerization.19-21 This
means that the dehydration of TBA (or the hydration
of isobutene) might also occur in the dimerization of
isobutene.
Although TBA dehydration and isobutene hydration
on ion-exchange resin catalysts have been studied
widely, few studies have been carried out under isobutene
dimerization conditions, i.e., in the liquid phase with

10.1021/ie049846s CCC: $27.50 2004 American Chemical Society


Published on Web 06/05/2004

Ind. Eng. Chem. Res., Vol. 43, No. 15, 2004 4061

low initial water content. The purpose of the present


work was to study the equilibrium of TBA dehydration
under these conditions and to construct a kinetic model
for the reaction. Experimental data were used to obtain
equilibrium constants and reaction enthalpies and to
determine parameters for kinetic equations derived
from various mechanistic assumptions.
Experimental Section
Reactor System. Experiments were carried out in
a stainless steel batch reactor (80 cm3) in the liquid
phase. The catalyst (1 g) was placed in the reactor in a
metal gauze basket. The reactor included a magnetic
stirrer and a mixing baffle. It was pressurized with
nitrogen to about 1.3 MPa to keep the reaction mixture
in the liquid phase, and it was held in a thermostated
water bath with which various temperatures (60-90 C)
could be maintained. The mixing speed in the batch
reactor was 1000 rpm so that no external mass-transfer
limitations would occur. Samples were taken manually
from the reactor via a sample valve.
Analytical Methods. The samples taken from the
reactor were analyzed with a Hewlett-Packard 5890
Series II gas chromatograph (GC) equipped with a DB-1
capillary column of length 60 m, film thickness 1.00 m,
and diameter 0.250 mm. A flame ionization detector was
used. The products were quantified by treating the
solvent as an internal standard.
Water in the product could not be analyzed by GC,
and the molar balance for isobutene was poor because
of its high volatility. The analysis of TBA can be
considered reliable, however, and the amounts of water
and isobutene were accordingly calculated from the TBA
molar balance.
Chemicals and Catalyst. TBA (Merck Schuchardt
OHG, >99%) was the reactant and isopentane (Fluka
Chemika AG, G99%) the solvent. Isooctane (Fluka
Chemika AG, G99.5%) was added to the reaction
mixture to study the reliability of the sampling. The
amount of water in the TBA was less than 0.1 wt %.
The catalyst was a commercial acidic ion-exchange
resin consisting of a styrene-divinylbenzene-based support to which sulfonic acid groups had been added as
active sites. The surface area of the dried catalyst was
measured to be 37 m2/g (BET analysis), the acid capacity
5.1 mmol/g (by titration22), and the particle size between
0.42 and 1.0 mm. Before use, the catalyst was dried
overnight in an oven at about 100 C. The water content
of the dried catalyst was measured to be 1.7 ( 0.5 wt %
(AMB 310 moisture analyzer).
Intraparticle Resistances. The dried catalyst was
sieved to different particle sizes and tested at 80 C in
a continuous flow system to study whether internal
diffusion limitations existed. The reactor system used
has been described in detail in our previous publication.21 The total flow of isopentane and 2 mol % of TBA
was about 36 g h-1. The conversion of TBA at the steady
state was the same (43-44%) in the experiments with
catalyst particle sizes of 0.42-0.59, 0.59-0.71, and
0.71-1.0 mm. Unfortunately, the experimental setup
did not allow for the testing of smaller particle sizes.
The results suggest that internal diffusion limitations
were not present. The unsieved catalyst was used in
further experiments.
Experiments. In this work, the dehydration of TBA
was studied in the absence of isobutene with TBA as
the sole reacting component in the initial reaction
mixture. Experiments were carried out with TBA con-

Figure 1. Mole fractions as a function of time in an experiment


at 60 C.

tents of 2-18 mol %, with an isooctane content 1 mol


%, and with isopentane as the solvent. The temperatures studied were 60-90 C, and the pressure was kept
at about 1.3 MPa. Several samples were taken from the
reactor during each experiment.
The reaction was carried out for about 6 h, during
which time the system reached equilibrium. Two or
more points were used in every experiment to determine
whether equilibrium was reached, and in uncertain
situations, the experiment was repeated. An example
of molar fractions as a function of time in a typical
experiment is presented in Figure 1. Under the conditions applied, isobutene did not dimerize appreciably
(diisobutene content < 0.06 mol % at 90 C). The
isobutene dimerization rate clearly decreases with
increasing TBA content.21 This explains the negligible
diisobutene formation in these experiments with relatively high TBA contents. Because of the low TBA
conversions, the isobutene and water contents were low
and only one liquid phase was observed.
Thermodynamics. Equilibrium constants Ka can be
calculated with equation

Ka )

aIBaH2O
aTBA

xIBxH2O IBH2O
xTBA

(1)

TBA

where ai is the activity of component i, xi is the


corresponding mole fraction at equilibrium, and i is the
activity coefficient. Activity coefficients were calculated
by the Dortmund modified UNIFAC method.23
Equilibrium constants can also be expressed in terms
of the Gibbs free energy change for the reaction (rG)

Ka ) exp -

rG
rH rS
+
) exp RT
RT
R

(2)

where rH is the enthalpy change and rS the entropy


change for the reaction and R is the universal gas
constant. If the temperature range investigated is
narrow, rH and rS can be assumed to be independent
of temperature.
Kinetic Modeling. The parameters of the kinetic
models were determined using Kinfit software24 with
the Levenberg-Marquardt optimization algorithm. In
the optimization, various kinetic models were combined
with an ideal batch reactor model, and the calculated
compositions were compared with the measured ones.
The temperature dependence of the rate constants
was described by the Arrhenius equation

k ) F exp -

E
RT

(3)

4062 Ind. Eng. Chem. Res., Vol. 43, No. 15, 2004
Table 1. Reaction Rate Equations
model

rate equation

-rTBA )

k(KaaTBA - aH2OaIB)

aIB +

-rTBA )

KH2O
KTBA
aTBA +
a
KIB
KIB H2O

-rTBA )

LH (two sites)

k(KaaTBA - aH2OaIB)
aTBA +

mechanism

KH 2O
KTBA

LH (one site)

aH2O

k(KaaTBA - aH2OaIB)
1 + KH2OaH2O

Petrus et al.26,27

where F is the preexponential factor, E is the activation


energy, and R is the universal gas constant. This
equation was reparametrized to the form

[ (

k ) Fref exp -

E 1
1
R T Tref

)]

TBA + H+ S TBAH+
TBAH+ S H2OH+ + IB
H2OH+ S H2O + H+

(4)

where Tref is the reference temperature and Fref and E


are the parameters to be optimized. Tref was chosen to
be 343 K (70 C). The adsorption equilibrium parameters were assumed to be independent of temperature.
In the tested models, the reaction on the surface of
the catalyst was considered as the rate-determining
step, and the active sites of the catalyst were assumed
to be equivalent. Parameters for a model that took into
account the different active sites that result from alcohol
adsorption9 were determined, but they did not give
satisfactory results.
Furthermore, adsorbed components were assumed to
occupy one surface site, and the reaction was assumed
to proceed through carbonium ions. Because very low
adsorption equilibrium constants (<1 10-8) were
obtained for isopentane in preliminary kinetic modeling
of the data, adsorption of the nonpolar components, i.e.,
isopentane and isooctane, was not included in the
models.
The reaction equilibrium constant Ka determined
experimentally was used in the models. The models are
presented in the following subsections, and the rate
equations for the dehydration of TBA can be found in
Table 1. The rates for the other components in the
dehydration of TBA follow the equation

rIB ) rH2O ) -rTBA

sites were unoccupied. Both types of models resulted


in very similar parameter values, suggesting that the
assumption made is justified. The model has four
parameters: the preexponential factor and activation
energy for the Arrhenius-type rate constant and the
adsorption equilibrium constants for TBA and water
calculated relative to the adsorption equilibrium constant of isobutene.
Model 2. LH1. Water is more polar than isobutene,
which means that it adsorbs preferentially on the active
sites.25 In the second model, the formed isobutene does
not adsorb on the catalyst, and only one active site is
needed. Thus, the second model follows a LangmuirHinshelwood-type mechanism with one active site.

(5)

Model 1. LH2. The first model that was tested follows


a Langmuir-Hinshelwood-type mechanism with two
active sites. This means that TBA and both products
(isobutene and water) adsorb on the active sites.

TBA + H+ S TBAH+
TBAH+ + H+ S H2OH+ + IBH+
H2OH+ S H2O + H+
IBH+ S IB + H+
Because the reaction takes place in the liquid phase, it
is assumed that there are no unoccupied active sites on
the catalyst. This assumption was verified by optimizing
the parameters for models in which some of the active

In this model, too, it was assumed that there are no


unoccupied active sites on the catalyst. Again, the model
in which some of the active sites were unoccupied was
tested, and it resulted in very similar parameter values.
For this model, three parameters have to be determined: the two parameters for the rate constant and
the adsorption equilibrium constant for water relative
to that of TBA.
Model 3. The third model is based on the mechanism
originally introduced by Petrus et al. for the hydration
of propene26 and linear olefins27 and also discussed by
Velo et al.13 in a study of isobutene hydration. In the
dehydration of TBA, this reaction mechanism follows
the equations

TBA + H+ S IBH+ + H2O


IBH+ S IB + H+
In deriving the rate equations, the steady-state hypothesis was used for the carbonium ion (IBH+). The
parameters include two parameters for the rate constant
and the adsorption equilibrium constant for water.
Results and Discussion
Thermodynamics. The equilibrium constants for the
reaction were calculated using eq 1 with the molar
contents in the equilibrium and activity coefficients
calculated by the Dortmund modified UNIFAC method.
The activity coefficients varied between 1 and 1.1 for
isobutene and isopentane, between 13 and 42 for water,
and between 2 and 8 for TBA.
Reaction equilibrium constants with different initial
TBA contents should be the same at the same temperature. Figure 2 presents the equilibrium constants at
60 C as a function of the initial TBA content. As can
be seen, with TBA contents higher than 5 mol %, the
equilibrium constant is between 0.17 and 0.20, but with
2 mol % TBA content, the equilibrium constant is about
0.3. This higher value is probably due to the activity
model, which does not predict the activity coefficient of
water correctly at very low concentrations. At low
concentrations, the activity coefficients are the highest.
Still, the activity model employed (Dortmund modified
UNIFAC) works better than the original UNIFAC model
at all conditions. The equilibrium constants that were

Ind. Eng. Chem. Res., Vol. 43, No. 15, 2004 4063
Table 2. Comparison of Reaction Enthalpies for the
Dehydration of TBA

ref

rH (kJ mol-1)

this work
Delion et al.12
Velo et al.13
Iborra et al.28

26
27.3-33.5a
26.5
39

Depending on the solvent.

Table 3. Estimated Parameter Values, Their Confidence


Limits and the Residual Sum of Squares (RSS)
parametera
Figure 2. Equilibrium constants Ka as a function of the initial
TBA content at 60 C.

Fref, mol/(s kgcat)


E, kJ/mol
KTBA/KIB
KH2O/KIB

model 1
0.08 ( 0.22
24 ( 7
0.77 ( 1.11
0.89 ( 1.18

Fref, mol/(s kgcat)


E, kJ/mol
KH2O/KTBA

model 2
0.21 ( 0.10
18 ( 6
1.5 ( 1.1

Fref, mol/(s kgcat)


E, kJ/mol
KH2O

model 3
13 ( 190
27 ( 6
130 ( 2000

Figure 3. Equilibrium constants Ka for the dehydration of TBA


as a function of temperature and corresponding values found in
the literature.12,28

calculated for experiments with initial TBA contents


under 4 mol % were accordingly assumed to be misleading and were omitted in calculating the average equilibrium constants.
The average equilibrium constants were calculated
using the results from four or more experiments at each
temperature. These constants as a function of the
inverse temperature are presented in Figure 3, which
shows that the equilibrium constant increases with
temperature. This indicates that the reaction enthalpy
is positive and that the dehydration of TBA is an
endothermic reaction. To obtain rH and rS, the
logarithm of Ka is presented as a function of the
temperature according to eq 2. With regression analysis,
we obtain

1
+ 7.6391
ln Ka ) -3111.9
T/K

(6)

The correlation constant R2 for the regression is 0.972.


From these results, we obtain a reaction enthalpy of 26
( 9 kJ mol-1 and a reaction entropy of 60 ( 30 J mol-1
K-1.
Figure 3 compares the obtained equilibrium constants
with those determined by Delion et al.12 using different
solvents and by Iborra et al.28 (both originally for
isobutene hydration). Our equilibrium constants are
almost the same as those determined by Delion et al.
with nitromethane solvent, whereas the constants obtained by Delion et al. with p-dioxane as the solvent and
those obtained by Iborra et al. are about 50% smaller.
In addition, Table 2 presents the reaction enthalpies for
the dehydration of TBA determined in this work along
with those obtained by Delion et al.,12 Velo et al.,13 and
Iborra et al.28 Iborra et al. obtained a reaction enthalpy
of 39 kJ mol-1 in their study of the simultaneous
syntheses of MTBE and TBA. The other enthalpy values
vary between 26 and 34 kJ mol-1, and thus, our value

value ( confidence limits

RSS
0.000 600

0.000 737

0.000 755

Parameters FRef and E refer to eq 4.

is of the same order of magnitude as the values reported


in these previous publications.
The water contents of the samples could not be
measured, and the calculations described above were
carried out by assuming that the amounts of isobutene
and water formed were equal. The effect of extra water
in the feed (the water content in TBA is <0.1 wt %, and
that in the catalyst is <2.2 wt %) was studied by adding
this maximum amount of additional water to the
equilibrium contents obtained earlier and by calculating
the new equilibrium constants. The change was less
than 2.1% at 60-80 C and 4.6% at 90 C. It should be
noted that the water content in TBA is probably less
than 0.1 wt %. Furthermore, the water in the catalyst
that desorbed at temperatures above 100 C during the
water content analysis probably does not participate in
the reaction.
Kinetic Modeling. Dynamic data from 23 experiments at four temperatures were used in the kinetic
modeling. The data with high error (>10%) in the
isooctane balance were not employed. Table 3 presents
the parameter values determined, the 95% confidence
limits, and the residual sum of squares for the models.
In model 1, only the activation energy is well identified, the value being 24 ( 7 kJ mol-1. The relative
adsorption equilibrium constants for water (KH2O/KIB)
and for TBA (KTBA/KIB) are correlated with each other
and with the preexponential factor (Fref, correlation
factor in the correlation matrix > 0.98). The other
correlations are weak (<0.6).
The parameters are better identified in model 2. The
activation energy (18 ( 6 kJ mol-1) is lower than that
obtained with model 1. Only the relative adsorption
equilibrium constant for water (KH2O/KTBA) and the
preexponential factor (Fref) are correlated.
The parameters for model 3 are not well identified,
and only the activation energy (27 ( 6 kJ mol-1) has
good confidence limits. High correlation exists between
the adsorption equilibrium constant for water (KH2O) and
the preexponential factor (Fref).
All three models (Table 1) give reasonable fits to the

4064 Ind. Eng. Chem. Res., Vol. 43, No. 15, 2004

The RSS values for models 2 and 3 are almost the


same. Because the parameters were better identified for
model 2, however, this model can be considered more
reliable. According to model 2, the dehydration of TBA
has an activation energy of 18 ( 6 kJ mol-1, and the
adsorption equilibrium constant of water is 1.5 times
that of TBA.
Conclusions
The thermodynamics of the dehydration of TBA was
studied, and the temperature-dependent equilibrium
constant that was obtained gave a reaction enthalpy of
26 kJ mol-1. The thermodynamic values agree well with
those reported in earlier studies on similar systems.
The kinetics of the dehydration of TBA was studied
as well, and parameters for three different models were
estimated on the basis of the measured data. The
Langmuir-Hinshelwood-type model with adsorption of
isobutene included did not give physically meaningful parameters because it gave higher adsorption equilibrium constants for isobutene than for water and
TBA. The other two models gave physically meaningful parameters as well as similar fits to each other.
The parameters were well identified for a LangmuirHinshelwood-type model in which isobutene adsorption
was omitted. The activation energy in this case was 18
kJ mol-1.
Acknowledgment
Funding from Fortum Oil and Gas Oy is gratefully
acknowledged.
Notation

Figure 4. Calculated amount of TBA (moles) as a function of


measured amount (moles).

data. This is evident from Figure 4, which shows the


calculated amounts of TBA as a function of the measured amounts for the three models. The closer the
points in the figure are to the diagonal, the better the
fit is. The models show similar trends, which means that
they present the data equally well.
Using model 1, the adsorption equilibrium coefficient
for isobutene was determined to be larger than the
coefficients for TBA and water. This does not seem
reasonable given that both TBA and water, as polar
components, should adsorb on the active sites more
readily than isobutene does.25 Because the parameter
estimation does not give physically meaningful parameters, the first model was discarded.
Because of its high activity coefficient, the activity of
water is higher than that of TBA and isobutene. The
adsorption of water on the catalyst is also stronger than
that of the other components.25 As a result, the term
KH2OaH2O dominates in the denominator of the rate
equations. This means that the ratio of the preexponential factor (Fref) to the adsorption equilibrium constant for water (KH2O), Fref/KH2O, can be considered as a
combined parameter and the parameters cannot be
determined separately. The high correlation between
the preexponential factors and adsorption equilibrium
constants observed in the models follows.

i ) activity coefficient of component i


ai ) activity of component i
E ) activation energy of the Arrhenius equation, kJ mol-1
F ) preexponential factor of the Arrhenius equation (eq
3), mol s-1 kgcat-1
Fref ) preexponential factor of reparametrized Arrhenius
equation (eq 4), mol s-1 kg cat-1
rG ) Gibbs free energy change for reaction, kJ mol-1
rH ) enthalpy change for reaction, kJ mol-1
k ) reaction rate constant, mol s-1 kgcat-1
Ka ) reaction equilibrium constant based on activities
Ki ) adsorption equilibrium constant of component i
R ) universal gas constant, 8.314 J mol-1 K-1
ri ) reaction rate of component i, mol s-1 kgcat-1
rS ) entropy change for reaction, J mol-1 K-1
T ) temperature, K
Tref ) reference temperature of reparametrized Arrhenius
equation (eq 4), K
xi ) molar fraction of component i
Abbreviations
DIB ) diisobutenes, 2,4,4-trimethyl pentenes
H+ ) proton of the catalyst
IB ) isobutene, 2-methyl propene
i-oct ) isooctane, 2,2,4-trimethyl pentane
IP ) isopentane, 2-methyl butane
LH ) Langmuir-Hinshelwood
MTBE ) methyl tert-butyl ether, 2-methoxy-2-methyl
propane
RSS ) residual sum of squares
TBA ) tert-butyl alcohol, 2-methyl-2-propanol

Literature Cited
(1) Beuther, H.; Kobylinski, T. P. The Chemistry of Oxygenates
Suitable for Use in Gasoline. Am. Chem. Soc. Div. Pet. Chem. 1982,
27, 880.

Ind. Eng. Chem. Res., Vol. 43, No. 15, 2004 4065
(2) OSullivan, D. A. Arco Increasing Chemicals Stake in
Western Europe. Chem. Eng. News 1985, 63 (11), 10.
(3) Abraham, O. C.; Prescott, G. F. Make Isobutylene from TBA.
Hydrocarbon Process. 1992, 71 (2), 51.
(4) Morse, P. M. Producers Brace for MTBE Phaseout. Chem.
Eng. News 1999, 77 (15), 26.
(5) Matouq, M. H.; Goto, S. Kinetics of Liquid-Phase Synthesis
of Methyl tert-Butyl Ether from tert-Butyl Alcohol and Methanol
Catalyzed by Ion Exchange Resin. Int. J. Chem. Kinet. 1993, 25,
825.
(6) Yin, X.; Yang, B.; Goto, S. Kinetics of Liquid-Phase Synthesis of Ethyl tert-Butyl Ether from tert-Butyl Alcohol and
Ethanol Catalyzed by Ion Exchange Resin and Heteropoly Acid.
Int. J. Chem. Kinet. 1995, 27, 1065.
(7) Assabumrungrat, S.; Kiatkittipong, W.; Sevitoon, N.; Praserthdam, P.; Goto, S. Kinetics of Liquid-Phase Synthesis of Ethyl
tert-Butyl Ether from tert-Butyl Alcohol and Ethanol Catalyzed
by -Zeolite Supported on Monolith. Int. J. Chem. Kinet. 2002,
34, 292.
(8) Gates, B. C.; Wisnouskas, J. S.; Heath, H. W., Jr. The
Dehydration of tert-Butyl Alcohol Catalyzed by Sulfonic Acid Resin.
J. Catal. 1972, 24, 320.
(9) Gates, B. C.; Rodriguez, W. General and Specific Acid
Catalysis in Sulfonic Acid Resin. J. Catal. 1973, 31, 27.
(10) Abella, L. C.; Gaspillo, P.-A. D.; Maeda, M.; Goto, S. Kinetic
Study on the Dehydration of tert-Butyl Alcohol Catalyzed by Ion
Exchange Resins. Int. J. Chem. Kinet. 1999, 31, 854.
(11) Gupta, V. P.; Douglas, W. J. M. Diffusion and Chemical
Reaction in Isobutylene Hydration within Cation Exchange Resin.
AIChE J. 1967, 13, 883.
(12) Delion, A.; Torck, B.; Hellin, M. Equilibrium Constant for
the Liquid-Phase Hydration of Isobutylene over Ion-Exchange
Resin. Ind. Eng. Chem. Process. Des. Dev. 1986, 25, 889.
(13) Velo, E.; Puigjaner, L.; Recasens, F. Inhibition by Product
in the Liquid-Phase Hydration of Isobutene to tert-Butyl Alcohol:
Kinetics and Equilibrium Studies. Ind. Eng. Chem. Res. 1988, 27,
2224.
(14) Velo, E.; Puigjaner, L.; Recasens, F. Intraparticle Mass
Transfer in the Liquid-Phase Hydration of Isobutene: Effects of
Liquid Viscosity and Excess Product. Ind. Eng. Chem. Res. 1990,
29, 1485.
(15) Heath, H. W., Jr.; Gates, B. C. Mass Transport and
Reaction in Sulfonic Acid Resin Catalyst: The Dehydration of tertButyl Alcohol. AIChE J. 1972, 18, 321.
(16) Ihm, S.-K.; Chung, M.-J.; Park, K.-Y. Activity Difference
between the Internal and External Sulfonic Groups of Macrore-

ticular Ion-Exchange Resin Catalysts in Isobutylene Hydration.


Ind. Eng. Chem. Res. 1988, 27, 41.
(17) Leung, P.; Zorrilla, C.; Recasens, F.; Smith, J. M. Hydration
of Isobutene in Liquid-Full and Trickle-Bed Reactors. AIChE J.
1986, 32, 1839.
(18) Leung, P. C.; Recasens, F.; Smith, J. M. Hydration of
Isobutene in a Trickle-Bed Reactor: Wetting Efficiency and Mass
Trasfer. AIChE J. 1987, 33, 996.
(19) Di Girolamo, M.; Lami, M.; Marchionna, M.; Pescarollo,
E.; Tagliabue, L.; Ancillotti, F. Liquid-Phase Etherification/Dimerization of Isobutene over Sulfonic Acid Resins. Ind. Eng. Chem.
Res. 1997, 36, 4452.
(20) Sloan, H. D.; Birkhoff, R.; Gilbert, M. F.; Nurminen, M.;
Pyhalahti, A. Isooctane Production from C4s as an Alternative to
MTBE. Presented at the NPRA 2000 Annual Meeting (AM-0034), San Antonio, TX, Mar 26-28, 2000.
(21) Honkela, M. L.; Krause, A. O. I. Influence of Polar
Components in the Dimerization of Isobutene. Catal. Lett. 2003,
87, 113.
(22) Fisher, S.; Kunin, R. Routine Exchange Capacity Determinations of Ion Exchange Resins. Anal. Chem. 1955, 27, 1191.
(23) Lohmann, J.; Joh, R.; Gmehling, J. From UNIFAC to
Modified UNIFAC (Dortmund). Ind. Eng. Chem. Res. 2001, 40,
957.
(24) Aittamaa, J.; Keskinen, K. I. Kinfit; Laboratory of Chemical Engineering, Helsinki University of Technology: Espoo, Finland, 2001.
(25) Helfferich, F. Ion Exchange; McGraw-Hill Book Co.: New
York, 1962.
(26) Petrus, L.; De Roo, R. W.; Stamhuis, E. J.; Joosten, G. E.
H. Kinetics and Equilibria of the Hydration of Propene over a
Strong Acid Ion Exchange Resin as Catalyst. Chem. Eng. Sci.
1984, 39, 433.
(27) Petrus, L.; De Roo, R. W.; Stamhuis, E. J.; Joosten, G. E.
H. Kinetics and Equilibria of the Hydration of Linear Butenes over
a Strong Acid Ion-Exchange Resin as Catalyst. Chem. Eng. Sci.
1986, 41, 217.
(28) Iborra, M.; Tejero, J.; El-Fassi, M. B.; Cunill, F.; Izquierdo,
J. F.; Fite, C. Experimental Study of the Liquid-Phase Simultaneous Syntheses of Methyl tert-Butyl Ether (MTBE) and tert-Butyl
Alcohol (TBA). Ind. Eng. Chem. Res. 2002, 41, 5359.

Received for review February 25, 2004


Revised manuscript received April 28, 2004
Accepted May 4, 2004
IE049846S

Você também pode gostar