Reaction From Dimethyl Carbonate DMC To

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

9862 Ind. Eng. Chem. Res.

2008, 47, 9862–9870

Reaction from Dimethyl Carbonate (DMC) to Diphenyl Carbonate (DPC). 2.


Kinetics of the Reactions from DMC via Methyl Phenyl Carbonate to DPC
J. Haubrock,*,† W. Wermink,† G. F. Versteeg,†,‡ H. A. Kooijman,‡ R. Taylor,‡
M. van Sint Annaland,† and J. A. Hogendoorn†
Department of Science and Technology, UniVersity of Twente, P.O. Box 217, 7500 AE Enschede,
The Netherlands, and Department of Chemical and Biomolecular Engineering, Clarkson UniVersity,
Potsdam, New York 13699-5705

The kinetics of the reaction of dimethyl carbonate (DMC) and phenol to methyl phenyl carbonate (MPC) and
the subsequent disproportion and transesterification reaction of methyl phenyl carbonate (MPC) to diphenyl
carbonate (DPC) have been studied. Experiments were carried out in a closed batch reactor in the temperature
range from 160 to 200 °C for initial reactant ratios of DMC/phenol from 0.25 to 3 and varying catalyst
(titanium-(n-butoxide)) concentrations. The concept of a closed, ideally stirred, isothermal batch reactor
incorporating an activity based reaction rate model has been used to fit kinetic parameters to the experimental
data taking into account the catalyst concentration, the initial reactant ratio DMC/phenol and the temperature.

1. Introduction the conversion of DMC toward MPC. A reactive distillation


process to produce DPC would need to be operated as close to
Diphenyl carbonate is a precursor in the production of chemical equilibrium as possible in order to achieve the highest
polycarbonate (PC) which is widely employed as an engineering possible conversion of the reactants toward DPC. This, in turn,
plastic in various applications basic to the modern lifestyle: in requires the reactions to proceed sufficiently fast. To assess
electronic appliances, office equipment, and automobiles, for whether reactive distillation is an attractive process alternative
example. About 2.7 million tons of PC are produced annually, to improve the conversion of DMC toward MPC, it is essential
a figure that is expected to increase by 5-7% yearly up to at to know how fast chemical equilibrium can be achieved, as this
least 2010.1 determines the required residence time in the reaction zone and,
Traditionally, PC is produced using phosgene as an interme- hence, also the dimensions of the equipment.
diate. This process suffers from a number of drawbacks: a 4
In this paper, we present reaction rate data for the conversion
ton portion of phosgene is needed for the production of 10 tons
of DMC to DPC at different initial molar ratios of DMC/phenol
of PC; phosgene is very toxic, and when it is used, the formation
in the temperature range between 160 and 200 °C. In addition,
of undesired salts cannot be avoided; and the process uses 10
an activity-based reaction rate model is employed to model the
times as much solvent (on a weight basis) as PC produced. The
experimental data, with the reaction rate constants being the
solvent, methylene chloride, is suspected to be carcinogenic,
actual fit parameters. The required activity coefficients as applied
and it is soluble in water. This results in a large amount of
in the reaction rate model are estimated with the UNIFAC group
wastewater that has to be treated prior to discharge.2
contribution method.4
Many attempts have been made to overcome the disadvan-
tages of the phosgene-based process.3 The main focus has been
2. Reactions
a route that produces diphenyl carbonate (DPC) via dimethyl
carbonate (DMC), which then reacts further with bisphenol-A The synthesis of diphenyl carbonate (DPC) from dimethyl
to form PC. The critical step in this route is the synthesis of carbonate (DMC) and phenol takes place through the formation
DPC from DMC that takes place via a transesterification reaction of methyl phenyl carbonate (MPC) and can be catalyzed either
to methyl phenyl carbonate (MPC), usually followed by a by homogeneous or heterogeneous catalysts. The reaction of
disproportionation and/or transesterification step to DPC. How- DMC to DPC is a two-step reaction. The first step is the
ever, in a batch reactor with an equimolar feed, an equilibrium transesterification of DMC with phenol to the intermediate MPC
conversion to MPC of only about 3% can be expected. and methanol (see Scheme 1).2,5
Therefore, creative process engineering is required to success- For the second step two possible routes exist: the transes-
fully carry out the reaction of DMC to DPC on a commercial terification of MPC with phenol (see Scheme 2) and the
scale. disproportionation of two molecules of MPC yielding DPC and
To make viable the process from DMC to DPC, the DMC (Scheme 3).
conversion of DMC to the intermediate MPC has to be According to Ono,2 side reactions also may occur (see
substantially increased. As one of the reaction products, Scheme 4). For this kind of reaction system, anisole is the main
methanol in this case, is the most volatile component in the byproduct, which can be formed from DMC and phenol through
mixture, it might be attractive to use reactive distillation to methylation.
remove methanol directly from the reaction zone to enhance

* To whom correspondence should be addressed. Phone: +49-2365- 3. Catalysts


49-2261. Fax: +49-2365-49-802261. E-mail: j.haubrock@
alumnus.utwente.nl. Numerous catalysts are known to promote common transes-

University of Twente. terification reactions. Nevertheless, many of these catalysts are

Clarkson University. not suitable for the transesterification of DMC to DPC as they
10.1021/ie071176d CCC: $40.75  2008 American Chemical Society
Published on Web 11/14/2008
Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008 9863
Scheme 1. Transesterification 1 Scheme 3. Disproportionation

Scheme 4. Side Reaction Forming Anisole, Methanol, and


Scheme 2. Transesterification 2 Carbon Dioxide

In another study, Fuming et al.8 presented a newly developed


catalyst, namely Samarium-trifluromethanesulfonate (STFMS)
also catalyze the decarboxylation to anisole.2 First, it has to be that was compared to a tin-based, a titanium-based, an aluminum-
decided whether a heterogeneous or homogeneous catalyst is based and a zinc-based catalyst, respectively. The selectivity
preferred. of the new STFMS catalyst toward DPC was between 2% (with
In industrial use, heterogeneous catalysts usually are consid- respect to (wrt) the tin based catalysts) and 12% (wrt the zinc
ered to be more desirable than homogeneous catalysts because based catalyst) higher, respectively. The conversion that was
of the ease of separation and regeneration of the catalyst.6 Most achieved in 12 h with the new STFMS catalyst was 35%, which
of these catalysts are supported metal oxides, such as MoO3 on is close to the conversion that can be obtained with the zinc-
silica. Ono2 investigated different types of heterogeneous based catalyst (37%) and the titanium-based catalyst (31%). The
catalysts and showed that the selectivity of these catalysts for aluminum- and zinc-based catalysts exhibited conversions of
the reaction toward anisole is much higher than that of DMC of less than 20% within 12 h of reaction time. It is
homogeneous catalysts where only traces of anisole were worthwhile to mention that while employing the samarium-,
detected. As anisole formation should be avoided if at all aluminum-, and zinc-based catalysts, there was always a small
possible, it was decided to adopt a homogeneous catalyst in amount of anisole formed which corresponded to roughly 1%
this study. of the converted DMC.8 Furthermore, it was reported that the
Advantages of the use of homogeneous catalysts when doing DPC obtained when using the tin-based catalyst showed a
kinetic experiments are as follows: grayish color due to the contamination with tin; this is
undesirable if DPC is used as a precursor for polycarbonate
(1) Experimental results may be used directly to derive the
production. All catalysts except the STFMS and the tin-based
kinetics. Mass transfer issues (e.g., diffusion limitations to/or
catalyst tend to hydrolyze in aqueous media. Thus, these
inside the catalyst particles) might cloud the interpretation of
catalysts cannot be utilized in aqueous environments and contact
experiments with heterogeneous catalysts.
with air also should be avoided.
(2) For the chemical system in this study, homogeneous For this study, titanium-(n-butoxide) was used as catalyst.
catalysts are commercially available and hence no tailor-made The selection for this catalyst is based on different consider-
manufacture of a supported heterogeneous catalyst is necessary. ations. A tin-based catalyst (n-Bu2SnO) has been disregarded
(3) The reproducibility of the kinetic experiments is better as the DPC made with this catalyst exhibits a grayish color,8
when utilizing a homogeneous, commercially available catalyst, so that this catalyst does not seem suitable for the manufacture
as the grade of this kind of catalyst will change only marginally of polycarbonate without additional processing steps. AlCl3 and
compared to tailor-made heterogeneous catalysts. ZnCl2 have been disregarded as these catalysts are both
Shaikh and Sivaram7 studied the performance of various susceptible toward water and thus tend to hydrolyze in the
homogeneous catalysts for the reaction of DMC and phenol and presence of aqueous media.8
the subsequent disproportionation of the intermediate MPC. The To avoid the laboratory synthesis of a new catalyst and to
reaction was carried out under continuous removal of the eliminate the undesired formation of anisole, also the use of
methanol/DMC azeotrope, while the temperature of the reaction the samarium-trifluromethanesulfonate (STFMS) has been re-
vessel was increased gradually from 120 to 180 °C. For the jected.8 The conversion rate of DMC as well as the selectivity
three tin-based catalysts, the authors reported yields of 41% and toward MPC and DPC with the titanium-(n-butoxide) catalyst
44% for DPC and yields of MPC between 3% and 12% achieved is close to that of the STFMS catalyst and the tin-based catalyst,
in 24 h of reaction time. The two titanium-based catalysts and no noticeable anisole is formed.6 Therefore titanium-(n-
exhibited yields between 27% and 33% for DPC and yields of butoxide) seems to be the most suitable catalyst to promote the
6% for MPC also achieved in 24 h reaction time. reactions from DMC to DPC (see Schemes 1-3). All chemicals
9864 Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008

Figure 1. Experimental setupsclosed batch reactor with supply vessel containing DMC.

used were either water free or contained only traces of water time to achieve equilibrium typically varied between 60 min
(e.g., phenol), which did not seem to affect the catalytic activity for the lowest catalyst amount and 15 min for the highest catalyst
during the experiments. amount (chemical equilibrium was judged to be reached when
a plot of the concentration profile as a function of time leveled
4. Experimental Work off). Additional experiments with mole fractions xcat of around
The main ingredients of the equipment used for the experi- 1.5 × 10-3 have been carried out, which demonstrated that
ments, shown schematically in Figure 1, were a stainless steel chemical equilibrium could be achieved in less than 5 min. This
reactor of 200 mL and a storage vessel of 300 mL. For a detailed already indicates that the reaction can be quite fast, using
description of the setup and the operation procedure, the reader reasonable amounts of catalyst. Some prior studies10,11 have
is referred to part 1 of this series.9 reported reaction times of up to 20 h, but from our present
To determine the reaction rate constants k1-k3 (see eqs 6-8) perspective, such long times do not seem to be required.
from the experimental measurements, the question arises which To ascertain that the initial level of mixing did not influence
key components should be used as indicator of the progress of the measurements, experiments with two different stirrer speeds
the reaction. As the equilibrium conversion of the reactants (800 and 1600 rpm) have been carried out, while otherwise
phenol and DMC is less than 3%, it was decided not to use the maintaining the same experimental conditions. At both stirrer
concentration of either to determine the reaction rate parameters
speeds the measured conversion rates were identical within
as the relative error would be large. Instead, the concentrations
experimental uncertainty. We conclude that it is safe to assume
of the products, methanol, MPC, and DPC, respectively, were
initially chosen as the key components. Eventually the use of that mixing is sufficient to justify adopting an ideally mixed
methanol as key component also has been omitted as GC reactor in order to model this process (see below).
analysis of the reactant DMC showed that it contained always The catalyst, titanium-(n-butoxide), tends to hydrolyze in the
a small amount (<1 wt %) of methanol. Apart from that, presence of water. To study the occurrence of any unwanted
methanol is also the most volatile component and a small part hydrolysis of the catalyst, several experiments with varying
of it might evaporate to the gas phase (5% of the overall amount amounts of catalyst were carried out, while maintaining the same
at V-L equilibrium), which would require an additional conditions (180 °C, reactant ratio phenol/DMC ) 1, nitrogen
correction of the GC results of the liquid samples. For these atmosphere). No detectable traces of degradation products in
reasons, also methanol was disregarded as a key component the GC diagram were found, which seems to indicate that
and only MPC and DPC have been taken as the key components catalyst degradation does not occur or, if it does, only to a
for the determination of the reaction rate parameters. Neverthe- negligible extent.
less, the initial amount of methanol introduced with the reactant
DMC has been taken into account in the analysis of the All chemicals employed in the experiments were used as
experimental data. received from the supplier. Dimethyl carbonate (purity: 99+%)
Experiments were carried out at temperatures of 160, 180, was purchased from Alldrich, phenol (99+%) and tetra-(n-butyl
and 200 °C at different catalyst concentrations and DMC/phenol orthotitanate) (98+%) were acquired from Merck. The catalyst
reactant ratios. The catalyst mole fractions xcat added in the was stored over molecular sieves (type 4a) to prevent degrada-
experiments varied between 1.0 × 10-4 and 4.5 × 10-4. The tion due to moisture from air.
Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008 9865

Figure 2. Dependence of the catalyst concentration on the temporal Figure 4. Time to reach 95% equilibrium versus the reciprocal catalyst
evolution of the MPC concentration (T ) 180 °C, molar ratio DMC/phenol mole fraction.
) 1).
backward reaction rate of MPC with methanol is much lower
than the forward reaction rate of DMC with phenol.
Figure 4 shows the time required to reach 95% of equilibrium
as a function of the reciprocal amount of catalyst. At the highest
catalyst mole fraction employed in the experiments (3.94 ×
10-4), the MPC mole fraction reached 95% of the mole fraction
at chemical equilibrium after just 10 min (see Figure 4). This
time increases to about 60 min for the lowest catalyst mole
fraction (1.21 × 10-4). This lends additional support to the
notion that we can model this system with a linear dependence
of the reaction rate on the catalyst mole fraction, not only for
the initial phase of the conversion but over the entire time of
these experiments where the backward reaction also becomes
important.

6. Do Both Reactions, Transesterification 2 and the


Disproportionation, Contribute to the DPC Formation?
Figure 3. Dependence of the initial reaction rate of MPC on the catalyst
mole fraction at 180 °C and a DMC/phenol ratio of 1 (markers ) initial
slope of concentration vs time for MPC taken at low MPC yields; solid
To investigate whether or not the disproportionation reaction
line ) linear fit). proceeds sufficiently fast, experiments were carried out with
no reactive component other than MPC present. Solvents DMC
5. Effect of Catalyst Concentration and phenol were replaced by inert n-heptane. In the absence of
phenol, MPC can only react to DPC via the disproportionation
To investigate the extent to which the catalyst concentration reaction and not at all via the transesterification reaction. Thus,
influences the reaction rate, a series of experiments with if DPC is found within the usual time frame of an experiment,
relatively low catalyst mole fractions (1.0 × 10-4 and 2.5 × this should mean that the disproportionation reaction is able to
10-4) were carried out at identical initial concentrations of DMC proceed at an appreciable rate and must be considered in the
and phenol and at the same temperature (180 °C). The results analysis of the data. To determine the influence of n-heptane,
of these experiments are shown in Figure 2, which suggest that an experiment with an equimolar ratio of DMC/PhOH, catalyst,
the equilibrium concentration is reached in times that vary from and 50 mol% n-heptane was carried out. This experiment yielded
around 15 min at the higher concentrations of catalyst to the same results in terms of the chemical equilibrium constant
something just over an hour at the lower concentrations. It has andscorrected for the mole fractions of DMC and phenolsthe
also been confirmed by additional experiments at 180 °C that same reaction rate as in the experiments without added solvent.
in the absence of catalyst the MPC mole fraction reaches only We infer that it is likely that n-heptane does not influence the
6% of the equilibrium mole fraction after 2 h. reaction rate in these experiments.
The initial forward rate of reaction 1 can be obtained by The results of the experiments with MPC dissolved in
differentiation of the data in Figure 2. The resulting linear fit n-heptane indicate that DPC and DMC are formed at nearly
dxMPC/dt|t)0 ) 0.138xcat [s-1] derived from the slope of the fitted the same rate (deviation <10%). Furthermore, the amounts of
straight line shown in Figure 3, suggests that the initial rate of DPC and DMC created are identical within experimental
reaction is directly proportional to the amount of catalyst and accuracy which supports the presumption that DPC and DMC
that the initial reaction rate can be expressed in the form: dxMPC/ are formed from MPC in equimolar amounts via the dispro-
dt|t)0 ) xcatkxDMC
initial initial
xphenol. This kinetic expression is equivalent portionation reaction. Methanol is present only in trace amounts.
to assuming that an elementary irreversible reaction (reaction This suggests that the formation of DPC issunder the present
1) is taking place. At low yields of MPC (<1%), the equilibrium experimental conditionssnot taking place via the transesteri-
yield of MPC is only around 2%sit can be assumed that the fication of MPC and phenol.
9866 Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008

7. Reaction Kinetics and Modeling Table 1. Activity Based Equilibrium Constants of Reactions 1-39
reaction activity based equilibrium value Ka,i
For a simple well-mixed batch reactor, the material balances
for the five components in the liquid phase can be written as 1 ln Ka,1 ) -2702/T[K] + 0.175
follows: 2 ln Ka,2 ) -2331/T[K] - 2.59
3 ln Ka,3 ) ln(Ka,2/Ka,1)
dxMeOH
) R1 + R2 (1) of the reaction mechanism is required that is not available in
dt
dxDMC the open literature. On the basis of the satisfactory description
) R3 - R1 (2) of the chemical equilibrium found by Haubrock et al.9 by
dt assuming elementary reactions, it seems reasonable to make the
dxPhOH same assumption here. The relations for the activity based
) -R1 - R2 (3) equilibrium values Ka,i valid in the temperature range from
dt
160-200 °C determined by Haubrock et al.9 are reproduced in
dxMPC
) R1 - R2 - R3 (4) Table 1.
dt In order to use Ka,i values to predict the equilibrium
dxDPC composition, we need to know the activity coefficients, for
) R2 + R3 (5) which Haubrock et al.9 used the UNIFAC method. It proved to
dt
be necessary to introduce a new UNIFAC group, the carbonate
We propose that the rates of the three reactions (Schemes
group O-CO-O, which was not then part of the published
1-3) can be expressed in the following form:
UNIFAC database. The interaction parameters of this new

(
R1 ) k1xcat γPhOHxPhOHγDMCxDMC -
1
γ x γ x
Ka,1 MPC MPC MeOH MeOH ) OCOO group with other UNIFAC groups, which are of
importance for the system presented here, were fitted to VLE
data for phenol-DMC, methanol-DMC, methanol-diethyl
(6)
carbonate (DEC), alkanes-DMC/DEC, alcohols-DMC/DEC,

(
R2 ) k2xcat γPhOHxPhOHγMPCxMPC -
1
γ x γ x
Ka,2 DPC DPC MeOH MeOH ) and toluene-DMC. For complete details, the reader is referred
to the Ph.D. thesis of Haubrock.13
(7) The temperature dependence of the reaction rate constants

( )
k1, k2, and k3 is accounted for by using the Arrhenius equation
1
R3 ) k3xcat γMPC2xMPC2 - γ x γ x (8) (eq 13):
Ka,3 DMC DMC DPC DPC
In eqs 6-8, xcat denotes the molar amount of catalyst, ki the ki ) k0,i exp(-EA,i /(RgasT)) (13)
forward reaction rate constant of reaction i, xj is the mole fraction There are, therefore, six parameters that need to be fitted to
of species j, γj the activity coefficient of species j, and Ka,i is experimental data using eqs 1-13: the three pre-exponential
the corresponding activity based chemical equilibrium constant. factors k0, i and the three activation energies EA,i. Our approach
A nearly identical approach has been applied by Steyer and was to fit k1, k2, and k3 for each temperature. An Arrhenius plot
Sundmacher12 to describe the reaction rate of the esterification may then be employed to determine k0,i and EA,i. The similarity
of cyclohexene with formic acid and subsequent splitting of of the reactions forming MPC and DPC by transesterification
the ester yielding cyclohexanol. As shown in Figure 2, the time suggests that k1 and k2 will be in the same order of magnitude.
taken to achieve chemical equilibrium depends strongly on the The influence of the catalyst amount on the reaction rate is
catalyst mole fraction used in the individual experiments; it accounted for via the catalyst mole fraction xcat as a linear factor
seems therefore justified and necessary to account for the catalyst implemented in the reaction rate equations (eqs 6-8). However,
amount by the introduction of the catalyst mole fraction in the this might not be sufficient as it is likely that not only the activity
reaction rate equation. One should note that the results in Figure coefficients of the reactants and products change at different
2 do not imply that the catalyst activity coefficient can also be process conditions but also the activity coefficient of the catalyst
considered constant for all experiments (e.g., for different DMC/ (γcat). In eqs 6-8, it is implicitly assumed that the activity
phenol ratios). coefficient of the catalyst γcat is constant and, therefore,
The activity based equilibrium constants of reactions 1-3 independent of the liquid phase composition. If this assumption
are given in eqs 9-11, with the overall chemical equilibrium is not justified, the rate constants (k1, k2, and k3) will probably
coefficient defined by eq 12: vary with composition as the influence of the “nonconstant”
aMPCaMeOH catalyst activity coefficient is in this case lumped into the
Ka,1 ) (9) optimized rate constants. Hence, in the interpretation of the
aDMCaPhOH experiments, k1, k2, and k3 will be optimized for each experiment
aDPCaMeOH conducted at a specific DMC/phenol reactant ratio. The opti-
Ka,2 ) (10) mized values of k1, k2, and k3 will subsequently be compared
aMPCaPhOH
to the optimized results of experiments at other DMC/phenol
aDPCaDMC Ka,2 reactant ratios.
Ka,3 ) ) (11)
aMPC 2 Ka,1 The batch reactor model neglects mass transfer from the liquid
phase (where the reaction takes place) to the gas phase; even
aDPCaMeOH2 for the most volatile component methanol, exploratory VLE
Ka,ov ) ) Ka,1Ka,2 ) (Ka,1)2Ka,3 (12) calculations have indicated that at gas-liquid equilibrium about
aDMCaPhOH2
95% of the amount of methanol formed will remain in the liquid
In the formulation of the reaction equilibrium equations, it phase (volume ratio liquid/gas phase ) 3:1). For the less volatile
has been assumed that reactions 1-3 represent elementary components, this percentage is near 100%. It should be noted
reaction steps. To justify this assumption, detailed knowledge that in case of an open system, as for example in a reactive
Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008 9867

Figure 6. Reaction rate constants ki as a function of the initial reactant


Figure 5. Reaction rate constants ki as a function of the initial reactant ratio DMC/phenol (T ) 180 °C).
ratio DMC/phenol (T ) 160 °C).
distillation column, especially methanol would steadily evapo-
rate from the liquid phase and mass transfer to the gas phase
should be taken into account.

8. Estimation of Rate Constants


Physically meaningful estimates as well as the likely range
of values of the reaction rate constant k1 were estimated from
the initial slopes of the mole fraction-time curve of MPC
applying the following equation dxMPC/dt|t)0 ) xcatkxDMC initial initial
xphenol.
As already discussed, at low conversions (yield MPC ∼ 1%),
the reverse reaction is not important and the estimation of k1 is
straightforward. The MPC mole fractions used for the estimation
of k1 were not corrected by the amount of DPC formed from
MPC as the amount of MPC which reacts further to DPC is
less than 2 mol % under the conditions investigated here.
The reaction rate constants of the first and second transes-
terification reaction k1 and k2 as well as the reaction rate constant
of the disproportionation reaction k3 have been optimized by Figure 7. Reaction rate constants k3 as a function of the initial reactant
fitting the theoretically predicted temporal evolution of the mole ratio DMC/phenol.
fractions of MPC and DPC (eqs 4 and 5). The actual fitting of
the rate coefficients was carried out using the simulation
environment gProms. The objective function minimized by
gProms was

N 1
Φ ) ln(2π) + min
2 2 {∑ ∑ ∑ [
NE NVk NMkj

ln(σkjm2) +

]}
K)1 j)1 m)1

(χ̃kjm - χkjm)2
(14)
σkjm2

9. Effect of the Reactant Ratio DMC/Phenol


The results of the optimization are summarized for k1 and k2
in Figure 5 (160 °C) and Figure 6 (180 °C) and for k3 in Figure
7 (160 and 180 °C). The average deviations between the
experimental and predicted values of the mole fractions were
less than 10% for MPC and less than 15% for DPC in the 95%
Figure 8. MPC mole fraction as function of time for different initial DMC/
confidence interval, respectively. In the case of very low phenol reactant ratios (T ) 180 °C): markers experimental data; continuous
experimental MPC and DPC mole fractions, usually observed line, model.
at DMC/phenol reactant ratios larger than two, the deviation
between a single experimental and predicted mole fraction can determined with the same experiments exhibit a somewhat larger
amount at most up to 15% for MPC and 30% for DPC, deviation between the experimental and model predicted values
respectively. of the mole fractions. However, the description of the experi-
This clearly shows that the experimental set of data can be mental mole fractions of DPC is still good (see Figure 9) and
used reliably to determine k1 whereas the k2 and k3 values the larger uncertainty in k2 and k3 does not yield a very large
9868 Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008

scatter when the individually determined values of k2 and k3


are plotted against the DMC/phenol ratio (see Figures 5-7).
The larger uncertainty in k2 and k3 is probably due to the very
low amounts of DPC formed in our experiments.
To obtain even more accurate data for k2 and k3, it would be
necessary to carry out experiments under continuous removal
of methanol. Exploratory experiments with MPC as starting
component have shown that significantly larger amounts of DPC
can be achieved; the absence of methanol means that the
backward reaction of DPC with methanol to MPC is suppressed.
The setup used in this study (Figure 1) does not allow for a
“reactive-distillation-like” mode where methanol is evaporated
at a specified pressure or temperature, respectively. Moreover,
a mass transfer model would be needed, to interpret this kind
of experiments and that would require physical property data
(e.g., kL values, diffusion coefficients, etc.) that are not known.
For these reasons, experiments under continuous removal of
methanol were not included within the scope of this study. The
values of k2 and k3 determined in this study should, therefore, Figure 9. DPC mole fraction as function of time for different initial DMC/
phenol reactant ratios (T ) 180 °C): markers, experimental data; continuous
be handled with some care, if used for completely different line, model.
conditions.
Figures 5 and 6 suggest that a linear relationship between relationship of k2 on the DMC/phenol ratio seems to suggest
the initial reactant ratio DMC/phenol and the reaction rate this fact, it cannot be concluded beyond a reasonable doubt
coefficients k1 and k2, respectively, can be deduced. The reaction because of the larger uncertainty in the individual k2 values.
rate constant k3 of the disproportionation reaction is not
Nevertheless, a possible change of the catalyst activity with
influenced by the DMC/phenol ratio (Figure 7), and this might
a change of the DMC/phenol ratio is not unlikely, considering
be attributed to different reaction mechanisms of the dispro-
the interaction of the in situ formed Ti-catalyst with the reactant
portionation reaction and the transesterification reactions. The
phenol. Assuming that one or more of the four butoxide ligands
scatter of the k3 values is around (10% (see Figure 7), which
of the titanium-(n-butoxide) catalyst have been substituted by
corresponds to the experimental uncertainty of the DPC mole
phenol,14 the in situ formed titanium(phenoxide) catalyst is likely
fractions used for the fitting of the reaction constant k3.
to show a substantial interaction with phenol. As the activity
Figure 6 shows that the fitted reaction rate constants k1 and
coefficient of phenol changes with a change of the DMC/phenol
k2 (180 °C) have nearly the same value (within (15% on
ratio, the activity of the catalyst is likely to change accordingly.
average). The deviation between the k1 and k2 (160 °C) values
There is no detailed information in the literature on the activity
shown in Figure 5 is largerson average (35%swhich is mainly of the titanium(phenoxide) catalyst as a function of the
due to the comparatively large k2 value at a DMC/phenol ratio composition of the mixture, and the present experiments also
of 2. From Figures 5 and 6, it can be concluded that the values do not provide enough information to unambiguously establish
of the kinetic constants k1 and k2 are similar andsfor DMC/ any such dependence; thus, the influence of the reactant ratio
phenol ratios < 1snearly identical. Considering the very similar of DMC/phenol on the catalyst activity can only be hypoth-
reactions, it could have been expected that also the reaction esized. In order to establish the activity of the catalyst as a
mechanism is identical and the kinetic rate similar which is function of composition, an extensive study including various
supported by the nearly identical reaction rate constants. vapor-liquid-equilibrium (VLE) experiments with changing
In view of the fact that activity coefficients of all reactants DMC/phenol ratios and catalyst concentrations would have to
and products were included in the model, we would hope that be carried out to determine the interactions between the catalyst
“nearly” constant values of the reaction rate constants might and the various species in the system. Moreover, the interpreta-
have been expected on purely fundamental grounds. This is tion of this kind of experiments is complicated by the fact that
obviously not the case for k1 and k2 (see Figures 5 and 6). There the species in the system are chemically reacting.
seem to be two possible reasons: it might be that either one or Looking again at Figures 5-7 and comparing the values of
more of the activity coefficients are inaccurate or that the catalyst k3 to those of k1 and k2, it can be seen that the reaction rate
activity changes with the DMC/phenol ratio. A previous study constant k3 is 1 order of magnitude larger than the values for k1
on the equilibria of reactions 1-3 has shown that the activity and k2. This indicates that the disproportionation reaction is
coefficient of DMC, important in the proper determination of intrinsically faster than the two transesterification reactions; this
k1, might be prone to error because the equilibrium value of is in agreement with the literature.15 Since there is normally an
the first transesterification reaction showed some variation excess of phenol in these experiments, the formation rates of
((20%) with the DMC/phenol ratio. As depicted in Figures 5 DPC by the second transesterification and the disproportionation
and 6, k1 changes by approximately a factor of 3 and it does reaction, respectively, are of the same order of magnitude.
not seem likely that an inaccuracy in the activity coefficient However, in industrial processes and at high reactant conver-
can be held responsible for the shifting value of k1. sions, the concentration of phenol might be much lower and
Accordingly, the linear increase of k1 most probably has to MPC much higher than in this study, so that the disproportion-
be attributed to a change in the activity of the catalyst. In which ation reaction may be the main route of MPC to DPC; this
case, it is also likely that k2 will be affected in the same way as should be kept in mind when scaling up the process.
k1; the two transesterification reactions are similar and we would Typical results of some experiments at different reactant ratios
expect that both reactions should be affected to a comparable together with the accompanying theoretical predictions are given
extent by a change in the catalyst activity. Although the linear in Figures 8 and 9 using the individual fitted reaction rate
Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008 9869

Figure 10. Model prediction: MPC mole fraction as a function of time for
different initial reactant ratios of DMC/phenol at 180 °C (xcat ) 1.50 × Figure 11. Arrhenius plot. Markers indicate fitted ki values derived from
10-4). kinetic experiments (T ) 160-200 °C; 1:1 DMC/phenol ratio): transes-
terification 1 (squares) and transesterification 2 (triangles).
constants k1, k2, and k3 (Figures 5-7) and activity based
equilibrium constants Ka,i at nearly identical catalyst amounts
(8.4 × 10-5 < xcat < 1.7 × 10-5). It can be seen that the
experimental MPC and DPC mole fractions are well in line with
the model predictions when using the individual fitted reaction
rate constants k1, k2, and k3 for the appropriate reactant ratio
DMC/phenol. This suggests that the proposed reaction rate
model is well suited to reproduce the experimental results.
Since the experiments as depicted in Figures 8-9 were carried
out at slightly varying catalyst mole fractions (8.4 × 10-5 <
xcat < 1.7 × 10-5), the reaction rates determined from the
individual experiments cannot directly be compared to each
other. However, the individually fitted k1, k2, and k3 values as
given in Figures 5-7 can be used with the batch reactor model
to simulate the concentration-time profiles scaled to the same
catalyst mole fraction (eqs 4 and 5) thereby excluding the effect
of the catalyst amount on the reaction rate and making it possible
to investigate the corresponding reaction rates at various DMC/ Figure 12. Arrhenius plot. Markers indicate fitted k3 values from kinetic
phenol ratios. In Figure 10, the course of the MPC mole fraction experiments (T ) 160-200 °C; 1:1 DMC/phenol ratio).
versus time for different DMC/phenol ratios using a catalyst Table 2. Values of the Pre-exponential Factor k0,i and the Activation
amount of xcat ) 1.50 × 10-4 is shown and it can be seen that Energy EA,i (Derived from Regression Lines in Figures 11 and 12,
DMC rich reactant mixtures (DMC/phenol ratio > 1) have only Respectively)
a slight influence on the formation rate of MPC, whereas in reaction k0,i [s-1] EA,i [kJ mol-1]
phenol rich reactant mixtures (DMC/phenol ratios < 1), the
reaction rate of MPC slows down considerably. The time to trans 1 (i ) 1) 2.42 × 108 73.5
trans 2 (i ) 2) 6.61 × 106 59.9
reach equilibrium roughly doubles for phenol rich reactant disprop (i ) 3) 14.88
mixtures.
As the same amount of catalyst is used in the simulations In Figure 11, the Arrhenius plots of k1 and k2 and the
shown in Figure 10, the different reaction rates can either be corresponding fitted ki values in the temperature range between
attributed to changing activity coefficients of the involved 160 and 200 °C are shown. As already mentioned earlier, the
species or to a varying catalyst activity. The product of the two transesterification reactions seem to have the same reaction
activity coefficients of DMC and phenol, γDMCγPhOH, changes mechanism and the numerical values of the reaction rate
only about 10% over the entire range of employed DMC/phenol constants derived from the experiments are similar. Hence, it
ratios and can, therefore, not be held responsible for the slower could be expected that the temperature dependence of the two
reaction rates observed for reactant ratios less than 1. This reaction rate constants k1 and k2 (Figure 11) would yield similar
supports the hypothesis that a varying catalyst activity is indeed activation energies. The Arrhenius plot of k3 and the corre-
responsible for the change in the reaction rate. sponding fitted k3 values in the temperature range between 160
and 200 °C are depicted in Figure 12. The disproportion reaction
10. Effect of Temperature exhibits no significant temperature dependencesthe scatter
shown in Figure 12 is within the experimental error margin.
An Arrhenius plot (Figures 11 and 12) may be used to
determine the pre-exponential factor k0,i and the activation
11. Conclusion
energy EA,i of the three reactions (see eq 13) from linear
regression. The results of such a calculation are summarized in In this study, the reaction rate constants of the transesterifi-
Table 2. cation reaction of DMC with phenol yielding the intermediate
9870 Ind. Eng. Chem. Res., Vol. 47, No. 24, 2008

MPC, the reaction rate constants of the consecutive transesteri- NE ) number of experiments performed (-)
fication reaction of MPC with phenol, and the reaction rate NMk,j ) number of measurements in the jth mole fraction in the kth
constants of the disproportionation of MPC have been experi- experiment (-)
mentally determined in a batch reactor. NVk ) number of variables measured in the kth experiment (-)
The influence of the catalyst concentration (titanium-(n- Rgas ) ideal gas constant (kJ mol-1 K-1)
butoxide)) and the temperature on the reaction rate constants Ri ) reaction rate of reaction i (s-1)
in the temperature range between 160 and 200 °C has been T ) temperature (K)
investigated as well as the influence of the initial reactant ratio t ) time (min or s)
of DMC/phenol. The concept of a closed, ideally stirred, xj ) mole fraction of component j (-)
isothermal batch reactor incorporating an activity based reaction x̃k,j,m ) mth measured value of mole fraction j in experiment k (-)
rate model, has been used to fit the values of the three reaction xk,j,m ) mth predicted value of mole fraction j in experiment k (-)
rate constants k1, k2, and k3 to the experimental data. γj ) activity coefficient of component j
The numerical values of the fitted reaction rate constants k1 σk,j,m2 ) variance of the mth measurement of mole fraction j in
and k2 are found to be similar whereas the numerical value of experiment k (-)
k3, belonging to the disproportionation reaction, is about 1 order Indices
of magnitude larger. Moreover, it was shown that the reaction cat ) catalyst (-)
rate constants of the two transesterification reactions (k1 and i ) reaction i (-)
k2) are strongly influenced by the initial reactant ratio of DMC/ j ) component j (-)
phenol which was attributed to inaccuracies in the activity k ) experiment k (-)
coefficients and to a changing catalyst activity. Nevertheless, m ) measurement m (-)
the change of the reaction rate constants over the initial reactant
ratio of DMC/phenol by a factor of 3 is too large to be caused
Literature Cited
only by flawed activity coefficients. Therefore, it is likely that
the activity coefficient of the catalyst changes over the initial (1) Westervelt, R. Polycarbonate. Chem. Week 2006, 168, 27.
reactant ratio of DMC/phenol. However, at the moment this (2) Ono, Y. Catalysis in the Production and Reactions of Dimethyl
Carbonate: An Environmentally Friendly Building Block. Appl. Catal., A
can only be regarded as a hypothesis as no detailed information 1997, 155, 133.
of the catalyst activity is available. Additional VLE experiments (3) Kim, W.; Joshi, U.; Lee, J. Making Polycarbonates Without
should be carried out to determine the interactions between the Employing Phosgene: An Overview on Catalytic Chemistry of Intermediate
catalyst and the other involved species yielding the activity and Precursor Synthesis. Ind. Eng. Chem. Res. 2004, 43, 1897.
coefficient of the catalyst to confirm the aforementioned (4) Fredenslund, A.; Jones, R. L.; Prausnitz, J. M. Group-Contribution
Estimation of Activity-Coefficients in Nonideal Liquid-Mixtures. AIChE
hypothesis. J. 1975, 21, 1086.
Experiments have shown that it seems necessary to remove (5) Fu, Z.; Ono, Y. Two-Step Synthesis of Diphenyl Carbonate From
methanol from the reaction mixture for two reasons: First, the Dimethyl Carbonate and Phenol Using MoO3/SiO2 Catalysts. J. Mol. Catal.
removal of methanol increases the conversion of DMC and A: Chem. 1997, 118, 293.
(6) Kim, W.; Lee, J. A New Process for the Synthesis of Diphenyl
phenol thereby promoting the formation of the intermediate Carbonate From Dimethyl Carbonate and Phenol Over Heterogeneous
MPC via transesterification 1. Second, in the absence of Catalysts. Catal. Lett. 1999, 59, 83.
methanol the disproportionation of MPC will contribute to the (7) Shaikh, A. G.; Sivaram, S. Dialkyl and Diaryl Carbonates by
overall conversion of MPC to DPC as the backward reactions Carbonate Interchange Reaction With Dimethyl Carbonate. Ind. Eng. Chem.
of transesterification 1 and 2, respectively, are suppressed. Res. 1992, 31, 1167.
(8) Fuming, M.; Guangxing, L.; Jin, N.; Huibi, X. A Novel Catalyst for
Therefore, the removal of methanol is important to achieve a Transesterification of Dimethyl Carbonate With Phenol to Diphenyl
selectivity toward DPC that is viable for industrial processes. Carbonate-Samarium Trifluoromethanesulfonate. J. Mol. Catal. A: Chem.
Reactive distillation might be used on an industrial scale not 2002, 184, 465.
only to allow for higher conversions of the reactants but also (9) Haubrock, J.; Raspe, M.; Versteeg, G. F.; Kooijman, H. A.; Taylor,
R.; Hogendoorn, J. A. The Reaction From Dimethyl Carbonate to Diphenyl
for a higher selectivity toward the desired product DPC. It is Carbonate. 1. Experimental Determination of the Chemical Equilibria Ind.
expected that the correlations presented in this paper could be Eng. Chem. Res. 2008, 47, 9854.
used in the modeling of reactive distillation processes for the (10) Shaikh, A. G.; Sivaram, S. Organic carbonates. Chem. ReV. 1996,
industrial relevant system presented in this work. 96, 951.
(11) Niu, H.; Guo, H.; Yao, J.; Wang, Y.; Wang, G. Transesterification
of Dimethyl Carbonate and Phenol to Diphenyl Carbonate Catalyzed by
Acknowledgment Samarium Diiodide. J. Mol. Catal. A: Chem. 2006, 259, 292.
(12) Steyer, F.; Sundmacher, K. Cyclohexanol Production via Esterifi-
The authors gratefully acknowledge the financial support of cation of Cyclohexene With Formic Acid and Subsequent Hydration of the
Shell Global Solutions International B.V. Tim Nisbet and Kees Ester-Reaction Kinetics. Ind. Eng. Chem. Res. 2007, 46, 1099.
Vrouwenvelder, both of Shell Global Solutions, are acknowl- (13) Haubrock, J. The Process of Dimethyl Carbonate to Diphenyl
edged for fruitful discussions. Furthermore, we would like to Carbonate: Thermodynamics, Reaction Kinetics and Conceptual Process
Design. Ph.D. Dissertation, University of Twente, The Netherlands, 2007;
thank H. J. Moed for the construction of the equipment and M. http://purl.org/utwente/58404.
Raspe for her contributions to the experimental work. (14) Sibum, H.; Guether, V.; Roidl, O.; Habashi, F.; Wolf, H. Ullmann’s
Encyclopedia of Industrial Chemistry - Titanium, Titanium Alloys, and
Notation Titanium Compounds; Wiley: New York, 2000.
(15) Buysch, H. Ullmann’s Encyclopedia of Industrial Chemistry -
aj ) activity of component j (-) Carbonic Esters; Wiley: New York, 2000.
EA,i ) activation energy of reaction i (kJ mol-1)
k0,i ) pre-exponential factor of reaction i (s-1) ReceiVed for reView August 29, 2007
Ka,i ) activity based equilibrium coefficient of reaction i (-) ReVised manuscript receiVed August 10, 2008
Accepted August 11, 2008
ki ) reaction rate constant of reaction i (s-1)
N ) total number of measurements taken during all experiments (-) IE071176D

You might also like