Sie sind auf Seite 1von 5

Ind. Eng. Chem. Res.

2008, 47, 53135317

5313

Reaction Kinetics for the Heterogeneously Catalyzed Esterification of Succinic


Acid with Ethanol
Aspi K. Kolah, Navinchandra S. Asthana, Dung T. Vu, Carl T. Lira, and Dennis J. Miller*

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): July 9, 2008 | doi: 10.1021/ie0706616

Department of Chemical Engineering & Materials Science, Michigan State UniVersity,


East Lansing, Michigan 48824

The reaction kinetics of the reversible esterification reaction of succinic acid with ethanol to form monoethyl
and diethyl succinate are presented. The reaction was studied in batch isothermal experiments catalyzed by
macroporous Amberlyst-15 ion-exchange resin. Experimental data were obtained between 78 and 120 C at
different mole ratios of ethanol to succinic acid and at ion-exchange resin catalyst concentrations from 1 to
5 wt % of solution. Kinetic modeling was performed using a pseudohomogeneous mole fraction model which
acceptably fits the experimental data. The kinetic model is useful for the design and simulation of processes
such as reactive distillation for diethyl succinate formation.
1. Introduction
Organic acid esters, produced by the reaction of organic acids
and alcohols, can be entirely biorenewable or green chemicals
that replace petroleum-based solvents. Succinic acid (1,4butanedioic acid, herein SA) can be esterified with alcohols such
as ethanol and n-butanol through a series reaction to yield diethyl
succinate (DES) and di-n-butyl succinate. A schematic reaction
scheme for esterification of succinic acid with an alcohol is
shown in Figure 1.
In addition to having low toxicity and low vapor pressure,
succinate esters have exceptional solvent properties and thus
find commercial application as solvents and in consumer
products such as paint strippers. In addition, succinate esters
are intermediates in the production of poly butylene succinate
(PBS) polymers, a polyester composed of SA and 1,4-butanediol
having attractive properties for broad application in automobiles
and consumer goods. The 1,4-butanediol is produced by
hydrogenation of succinate ester,1,2 hence the entire PBS
polymer is a succinate based, biorenewable material. Esters of
SA (primarily dimethyl esters) are also being investigated for
their insulinotropic potential in rats.37
A conventional process for DES production involves a stirred
batch or continuous reactor with sulfuric acid as a homogeneous
catalyst. Because the extent of reaction is thermodynamically
limited, intermediate product removal and multiple reaction
stages are required to achieve complete SA conversion. Many
of the difficulties in using homogeneous catalysts can be
eliminated through the use of heterogeneous catalysts such as
ion exchange resins or supported clays. The heterogeneous
catalyst allows easy mechanical separation of the catalyst from
reaction media by decantation or filtration, reduces or eliminates
corrosion problems, and facilitates continuous process operation.
For reviews on the use of heterogeneous ion exchange resins
as catalysts prior to 1995, the reader is referred to the works of
Olah et al.,8 Chakrabarti and Sharma,9 and Sharma et al.10 For
a more recent review, the reader is referred to the work of
Harmer and Sun.11
Prior information in the literature on the kinetics of SA
esterification with ethanol or n-butanol is scarce. Saigo et al.12
have synthesized succinic acid esters using phosphinechalcogenide as a catalyst. Recently, Benedict et al.13 have described
* To whom correspondence should be addressed. Tel.: 1-517-3533928. Fax: 1-517-432-1105. E-mail: millerd@egr.msu.edu.

a process for the pervaporation-assisted esterification of lactic


acid and SA with downstream ester recovery using Amberlyst
XN-1010 and Nafion NR50 as catalyst.
At present no information is available in the open literature
describing the kinetics of SA esterification with ethanol in the
presence of ion-exchange resin catalysts. Thus, we have
conducted isothermal experimental batch studies on the esterification of SA with ethanol in the presence of Amberlyst-15
ion-exchange resin as catalyst. A pseudohomogeneous mole
fraction based kinetic model is presented for correlation of the
experimental data. This kinetic model is useful for designing
processes such as continuous SA esterification using reactive
distillation.
2. Experimental Details
2.1. Materials. For the kinetic experiments, anhydrous
succinic acid crystals (>99%), herein SA, were obtained from
Sigma-Aldrich. Absolute ethanol (>99%) and HPLC grade
water were obtained from J. T. Baker, Inc. The strong cation
exchange resin catalyst Amberlyst-15 (Rohm and Haas, Philadelphia, PA) was obtained in H+ form (dry, 0.6 mm beads)
and was used without modification; the resin acidity as specified
by the manufacturer is 4.6 equiv H+/kg dry resin. Purity of all
chemicals was checked by gas chromatography or HPLC.
2.2. Analysis. The presence of succinic acid (SA), monoethyl
succinate (MES), and diethyl succinate (DES) was first confirmed by GC-MS analysis of their trimethyl silyl (TMS)
derivatives. For reaction samples, SA and its ethyl esters (MES
and DES) were quantitatively analyzed on a Hewlett-Packard
1090 HPLC using a reversed phase C18 column (Novapak, 3.9
mm 150 mm) held at 40 C. Water/acetonitrile (ACN)
mixtures, buffered at pH of 1.3, were used as the mobile phase
(0.8 mL/min) in a gradient mode (0% ACN at t ) 0 min to
60% ACN at t ) 20 min to 90% ACN at t ) 25 min to 0%
ACN at t ) 28 min). The species were quantified by UV
detection (Hitachi L400H) at a wavelength of 210 nm. Succinic
acid and DES were identified and quantified by comparing
HPLC retention time and peak area with their respective
calibration standards. Pure standard for MES could not be
commercially obtained. On a mass basis, the response factor
for DES was found to be 1.11 times higher than that for SA;
therefore, the response factor for MES was calculated as an
average of response factors for SA and DES. Using this response

10.1021/ie0706616 CCC: $40.75 2008 American Chemical Society


Published on Web 07/09/2008

5314 Ind. Eng. Chem. Res., Vol. 47, No. 15, 2008

Figure 1. Esterification of succinic acid with an alcohol.

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): July 9, 2008 | doi: 10.1021/ie0706616

Table 1. Summary of Kinetic Studies and Average Prediction


Errors

run
no.

figure
no.

1
2
3
4
5
6
7
8
9
10
11
12

S1
2
S2
S3
S4
S5
S6
S9
S7
S8
S10
S11

initial
mole ratio
EtOH:SA

catalyst
loading
(wt %)

temp
(C)

average
error (%)
in succinate
species mole
fraction
(Frel, eq 14)

10:1
10:1
10:1
10:1
10:1
10:1
10:1
10:1
10:1
10:1
15:1
20:1

2
2
2
2
2
1
3
5
1
5
2
2

78
90
100
110
120
90
90
90
78
78
90
90

25
29
23
17
18
28
22
25
26
19
27
28

average error
(absolute) in
succinate
species mole
fraction
(Fabs, eq 13)
0.0048
0.0052
0.0035
0.0025
0.0025
0.0052
0.0029
0.0035
0.0057
0.0045
0.0028
0.0022

factor for MES, the carbon balance for each reaction sample,
based on carbon in the starting SA, was in the range of 100 (
10%.
Reaction samples were analyzed for water content using a
Varian 3700 gas chromatograph equipped with thermal conductivity detector (TCD) and a stainless steel column (4 m
3.25 mm) packed with a stationary phase of Porapak Q. The
column oven temperature was held at 413 K for two minutes,
then ramped at 20 K/min to 493 K, where it was held for 6
min. n-Butanol was used as an internal standard. High purity
helium (99.999%) was used as carrier gas at a flow rate of 20
mL/min. The injector and detectors were maintained at 493 K.
Samples were analyzed for ethanol and byproduct diethyl
ether (DEE) using a Perkin-Elmer Sigma-2000 gas chromatograph equipped with flame ionization detector (FID) and a
bonded-phase fused silica capillary column (SPB-5, 30 m
0.53 mm). The column oven was held at 313 K for 7 min,
ramped at 2 K/min to 473 K where it was held for 5 min.
Anhydrous toluene was used as an internal standard. High purity
helium (99.999%) was used as carrier gas at a flow rate of 10
mL/min. The injector and detectors were maintained at 493 K.
2.3. Batch Kinetic Experiments. Esterification reactions at
78 C were performed in a 2 10-4 m3 jacketed glass reactor
equipped with a recirculating constant temperature oil bath. The
reaction volume was 1.1 10-4 m3. A spiral coil condenser,
open to the atmosphere, was placed on top of the reactor. The
glass reactor was equipped with temperature and stirrer speed
monitoring devices and a sampling port. In operation, measured
quantities of ethanol and succinic acid were added to the reactor,
and heating and stirring were started simultaneously. Once the
desired temperature was achieved, usually in about 15 min,
catalyst (Amberlyst-15 ion-exchange resin) was added for the
case of resin catalyzed reactions and the stirring speed was
increased to 800 rpm. This point in time was considered as the
zero reaction time. Samples were withdrawn at specific time
intervals and immediately transferred to an ice bath (prior to
analysis) in order to ensure that no further reaction took place.
Esterification reactions above 78 C were performed in a 1
10-4 m3 stainless steel autoclave (5000 Multireactor System,
Parr Instrument Co.) equipped with temperature and stirrer speed

Figure 2. Catalytic esterification of succinic acid with ethanol. Reaction


conditions: mole ratio of ethanol to succinic acid, 10:1; Amberlyst 15 cation
exchange resin catalyst loading, 2 wt %; reaction temperature, 90 C. (9)
SA; (b) MES; (2) DES.

Figure 3. Initial rate of succinic acid esterification vs catalyst loading.


Reaction conditions: mole ratio of ethanol to succinic acid, 10:1. (() -90;
(9) -78 C.

monitors and a sampling port. In operation, measured quantities


of ethanol, SA, and catalyst for resin-catalyzed reactions were
added to the reactor and heating was started with slow stirring.
The total reaction volume was maintained between 5.5 10-5
to 6.0 10-5 m3. The desired temperature was achieved in
about 15 min, at which time the stirring rate was increased to
740 rpm. This time was considered as the zero reaction time.
Samples were withdrawn at specific time intervals through a
cooled metal tube and immediately transferred to an ice bath
in order to ensure no further reaction took place before analysis.
All samples were analyzed using the method described in section
2.2.
3. Results and Discussion
Batch kinetic experiments were carried out to study the effects
of reaction temperature, catalyst loading, and initial reactant
molar ratio on the cation-exchange resin catalyzed esterification
of SA with ethanol. It was observed from varying stirrer speed
in initial experiments that conversion rates were unaffected at
stirring speeds above 500 rpm, so all kinetic experiments were
performed at 800 rpm to avoid external mass transfer limitations.
Table 1 shows the reaction conditions for all of the experiments performed in this work. In Table 1, catalyst loading (wt
%) is based on mass of the liquid phase. Ethanol:SA molar ratios
>10:1 were required because of limited succinic acid solubility
(12 wt %) in ethanol at low temperature.
A set of time trajectories of SA, MES, and DES at typical
batch esterification conditions is shown in Figure 2. The species
time trajectories for all other experiments are presented as
Figures S1-S12 in the Supporting Information.

Ind. Eng. Chem. Res., Vol. 47, No. 15, 2008 5315

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): July 9, 2008 | doi: 10.1021/ie0706616

The dependence of the succinic acid conversion rate on


temperature over the range from 78 to 120 C is given in Figures
S1-S4 of the Supporting Information. The effect of catalyst
loading on initial SA esterification rate at an ethanol:SA molar
feed ratio of 10:1 is shown in Figure 3. The initial SA
esterification rate was determined by fitting the experimental
succinic acid mole fraction vs time with a polynomial equation
and then evaluating the derivative at t ) 0. The initial SA
esterification rate is proportional to catalyst loading, an indication that the observed reaction kinetics are independent of
external mass transport resistances. Additional experiments at
different catalyst loadings are given in Figures S5-S9. The
effect of the ethanol:SA initial reactant mole ratio was studied
at 90 C and 2% catalyst loading; results are given in Figures
S10 and S11.

rate constants ki are included in the effectiveness factor in the


model equations. The complete equations are solved to determine the values of kinetic constants that best describe the
reaction data.
4.3. Mole Fraction based Kinetic Model. On the basis of
the reactions in eqs 1 and 2, a pseudohomogeneous, mole
fraction based kinetic model for ion-exchange resin-catalyzed
SA esterification has been developed, and kinetic data from a
wide range of reaction conditions, including runs 1-12 in Table
1, have been used to determine kinetic constants for the
reactions. The rate of formation of each species in the reaction
mixture is described by eqs 59 below:
-

4.1. Kinetic Pathways. Reactions 1 and 2 below describe


the pathways involved in the esterification of SA with ethanol.
k1

SA + EtOH {\} MES + W

wcatk11

(1)

k1/Keqx,1
k2

MES + EtOH {\} DES + W

(2)
-

k2/Keqx,2

Reactions 1 and 2 are the series reactions to form DES from


SA via intermediate formation of MES (eq 1).
4.2. Mass Transfer Considerations. The influence of intraparticle mass transfer resistances on Amberlyst-catalyzed
esterification was evaluated by first calculating of the observable
modulus (2) and implementation of the Weisz-Prater
criterion (2 , 1) for each experiment. Taking the initial rate
as the maximum rate for each experiment, calculating the liquid
phase effective diffusivity of succinic acid in ethanol as 5.8
10-10 m2/s via the Wilke-Chang equation, and applying the
experimental observation (via volumetric measurement) that the
ion-exchange resin particle swells by 50% upon exposure to
ethanol to a particle diameter of 0.69 mm, the values of
observable modulus ranged from 0.3 at 78 C to 1.5 at 120 C.
The reaction is thus moderately mass transfer limited at the
temperature range investigated. Succinic acid esterification is
therefore somewhat more rapid than for other acids in similar
esterification studies.1416
Mass transfer limitations were accounted for by including
the intraparticle effectiveness factor in the kinetic model
developed herein. Although the esterification reactions are
bimolecular and reversible, the effectiveness factors are calculated assuming that reactions 1 and 2 (eqs 1 and 2) are
irreversible, first order, isothermal reactions. This is justified
because ethanol is always in considerable excess (>10:1 molar
ratio with succinic acid) and the equilibrium conversion in all
experiments exceeds 98% for reaction 1 (eq 1) and 82% for
reaction 2 (eq 2). The resulting equations for Thiele modulus
() and effectiveness factor () are given as
i ) (dp/6)(kiFcatxEtOH/Fsoln)0.5 i ) 1, 2

(3)

i ) tanh(i)/i i ) 1, 2

(4)

where the catalyst and solution densities in eq 3 are required to


place the rate constant on a per unit volume catalyst basis and
the ethanol mole fraction, taken as the average over the
experiment, reflects the pseudofirst order assumption in the
effectiveness factor calculation.
In the kinetic model presented in the next section, the
preexponential factors and activation energies for the forward

)
)

(5)

dxMES
xDESxW
) wcatk22 xMESxEtOH +
dt
Keqx,2

4. Kinetic Modeling

(
(

dxSA
xMESxW
) wcatk11 xSAxEtOH dt
Keqx,1

(
(

xMESxW
- xSAxEtOH
Keqx,1

)
)

dxDES
xDESxW
) wcatk22
- xMESxEtOH
dt
Keqx,2

dxEtOH
xMESxW
) wcatk11 xSAxEtOH +
dt
Keqx,1

wcatk22 xMESxEtOH -

xDESxW
Keqx,2

(7)

(8)

(9)

dxW
xMESxW
) wcatk11
- xSAxEtOH +
dt
Keqx,1

wcatk22
where

(6)

xDESxW
- xMESxEtOH
Keqx,2

( )

-EA,i
(10)
RT
4.4. Chemical Equilibrium Constants. The chemical equilibrium constants are given by
ki ) k0i exp

Keqx,i )

i
i,eq

(11)

The value of mole fraction equilibrium constants Keqx,i for


reactions under consideration were determined by analysis of
the experimental data at long reaction times (e.g., approaching
equilibrium) and were found to be 5.3 and 1.2 for the formation
of MES and DES, respectively. These constants were taken to
be independent of temperature.
4.5. Determination of Rate Constants. The kinetic eqs 59
were numerically integrated via a fourth order Runge-Kutta
method using ordinary differential equation solver ode23 in
Matlab 7.0 to determine optimum values of the four adjustable
kinetic parameters for the resin-catalyzed reactions, the preexponential factors k01 and k02, and the energies of activation EA,1
and EA,2. Starting with an initial set of rate constants, the liquid
phase mole fractions for all species over the course of reaction
were calculated and compared with the experimental values.
The four kinetic parameters were then incremented sequentially
in order to minimize the sum of the mean square differences
given by

Fmin2 )

(xj,cal - xj,expt)2

samples

nsamples

(12)

5316 Ind. Eng. Chem. Res., Vol. 47, No. 15, 2008

this value with TOF values for other organic acid esterification
reactions that we have investigated in our laboratory. At the
same conditions, the TOF for lactic acid esterification17 with
ethanol is 13 1/h; for citric acid,16 the TOF for the first
esterification step is 2.1 1/h. The similarity in TOF between
lactic acid and succinic acid is not surprising; however, the
significantly lower TOF for citric acid suggests steric hindrance
associated with multiple carboxylic acid functionalities on citric
acid or more generally a reduced access of the bulkier citric
acid molecule to the resin acid sites.

Table 2. Parameters for Resin-Catalyzed Succinic Acid


Esterification with Ethanol
parameter
0

k1
k20
EA,1
EA,2
Keqx,1
Keqx,2

units

value

kgsoln/kgcats
kgsoln/kgcats
kJ/kmol
kJ/kmol

5.3107
8.0107
66000
70000
5.3
1.2

Table 3. Effectiveness Factors for Succinic Acid Esterification


temperature (K)

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): July 9, 2008 | doi: 10.1021/ie0706616

351
363
373
383
393

1 eq 1

2 eq 2

0.94
0.88
0.80
0.70
0.60

0.97
0.95
0.90
0.84
0.76

5. Conclusions

After this optimization, the calculated mole fractions of the


succinate components (SA, MES, DES) were compared to the
corresponding experimental values to calculate an average
absolute error in mole fraction (Fabs) and a relative average error
in percent (Frel), giving the mean relative deviation, represented
both absolutely and on a percentage basis, shown below.

Fabs )

Frel )

|xj,cal - xj,expt|

samples

samples

nsamples
xj,cal - xj,expt
xj,expt

nsamples

(13)

Succinic acid esterification kinetics were studied at reaction


temperature from 78 to 120 C, ethanol:SA initial mole ratios
from 10:1 to 20:1, and Amberlyst 15 cation exchange resin
loadings from 1% to 5% of solution. Esterification kinetics are
described using a mole fraction based pseudohomogeneous
model that includes intraparticle mass transfer limitations
expressed as effectiveness factor for a first order irreversible
reaction. The rate expressions describe the kinetics of MES and
DES formation over a wide range of catalyst concentration,
reactant molar ratios, and temperature. The model presented can
be conveniently used for design and scale-up of integrated
processes like reactive distillation for synthesis of DES.
Acknowledgment
Financial support from the National Corn Growers Association and the U.S. Department of Energy are greatly appreciated.

100 %

(14)

The values of the kinetic parameters are shown in Table 2.


Predicted mole fractions are given as continuous lines in Figure
2 and Figures S1-S11 in the Supporting Information. It can be
observed from the Figures and from Table 2 that the model
predicts succinic acid esterification experimental data reasonably
well.
The values of the absolute mole fraction error Fabs and relative
mole fraction error Frel are reported in Table 1. For SA, the
average percent error in mole fraction is highest in the region
when the SA concentration is low. Large errors are also observed
in the case of DES in the initial reaction period up to about
300 min, where its concentration is low.
Values of the effectiveness factors for each reaction at each
temperature are given in Table 3. The reaction exhibits
significant mass transport resistances at temperatures above 100
C. It should be noted that we attempted to fit the kinetic data
without including intraparaticle mass transport limitationssthe
average errors were one to two percentage points higher than
when effectiveness factor was included.
For the reaction system under consideration, the formation
of diethyl ether from the etherification reaction of two molecules
of ethanol has not been included, since the esterification reaction
is very fast in comparison to the kinetics of diethyl ether
formation. The kinetics of diethyl ether formation over Amberlyst 15 resin have been reported in an earlier publication
from the authors laboratory.16
The turnover frequency (TOF, (mol SA/mol H+ sites/h)) for
SA to MES on Amberlyst 15 resin acid sites can be easily
determined from the acid site density (4.6 mol H+/kg resin),
catalyst loading (wcat), initial SA concentration (mol/kgsol), and
rate constant k1. At 90 C and an initial SA concentration of
1.0 mol/kgsol in ethanol, corresponding to a 19:1 EtOH:SA feed
ratio, the turnover frequency is 12/h. It is interesting to compare

Nomenclature and Units


DES ) diethyl succinate
dp ) Amberlyst 15 ion-exchange resin catalyst particle diameter;
1.1 10-4 m
EA,i ) energy of activation for reaction i (kJ/kmol), i ) 1, 2
EtOH ) ethanol
ki ) rate constant for catalyzed reaction i, i ) 1, 2 (kgsoln/kgcat s)
k0i ) pre-exponential factor for catalyzed reaction i, i ) 1, 2 (kgsoln/
kgcat s)
Keq,i ) mole fraction based reaction i equilibrium constant, i ) 1,
2
MES ) monoethyl succinate
R ) gas constant; 8.31 (kJ/kmol K)
SA ) succinic acid
T ) temperature (K)
W ) water
wcat ) catalyst concentration (kgcat/kgsoln)
xj ) mole fraction of jth component in liquid phase solution
i ) intraparticle effectiveness factor for reaction i (i ) 1, 2)
i ) Thiele modulus for reaction i (i ) 1, 2)
Fcat ) catalyst particle density; 1200 (kg/m3)
Fsoln ) reaction solution density; 880 (kg/m3)
Subscripts
i ) reaction index
j ) component in solution

Supporting Information Available: Graphs of species


concentration vs time for a wide range of reaction conditions.
This material is available free of charge via the Internet at
http://pubs.acs.org.
Literature Cited
(1) Turek, T.; Trimm, D. L. The Catalytic Hydrogenolysis of Esters to
Alcohols. Catal. ReV. Sci. Eng. 1994, 36 (4), 645.

Downloaded by INDIAN INST OF TECHNOLOGY BOMBAY on September 4, 2015 | http://pubs.acs.org


Publication Date (Web): July 9, 2008 | doi: 10.1021/ie0706616

Ind. Eng. Chem. Res., Vol. 47, No. 15, 2008 5317
(2) Attig, T. G.; Graham, A. M. Process and Catalysts for the
Manufacture of Gamma-Butyrolactone and 1,4-Butanediol by Hydrogenation
of Maleic Acid. U.S. Patent 4,827,001, 1989.
(3) Mukala-Nsengu, A.; Fernandez-Pascual, S.; Martn, F.; Martn-delRo, R.; Tamarit-Rodriguez, J. Similar Effects of Succinic Acid Dimethyl
Ester and Glucose on Islet Calcium Oscillations and Insulin Release.
Biochem. Pharmacol. 2004, 67, 981.
(4) Laghmich, A.; Ladrie`re, L.; Dannacher, H.; Bjorjling, F.; Malaisse,
W. J. New Esters of Succinic Acid and Mixed Molecules Formed by Such
Esters and a Meglitinide Analog. Study of Their Insulinotropic Potential.
Pharmacol. Res. 2000, 41, 543.
(5) Kadiata, M. N.; Malaisse, W. J. Opposite Effects of D-Glucose
Pentaacetate and D-Galactose Pentaacetate Anomers on Insulin Release
Evoked by Succinic Acid Dimethyl Ester in Rat Pancreatic Islets. Life Sci.
1999, 64, 751.
(6) Ladrie`re, L.; Malaisse-Lagae, F.; Verbruggen, I.; Willem, R.;
Malaisse, W. J. Effects of Starvation and Diabetes on the Metabolism of
[2,3-13C] Succinic Acid Dimethyl Ester in Rat Hepatocytes. Metabolism
1999, 48, 102.
(7) Leclercq-Meyer, V.; Malaisse, W. J. Potentiation of Glucagon-Like
Peptide 1- Insulinotropic Action by Succinic Acid Dimethyl Ester. Life Sci.
1996, 58, 1195.
(8) Olah, G. A.; Iyer, P. S.; Prakash, G. K. S. Synthesis 1986, 513.
(9) Chakrabarti, A.; Sharma, M. M. Cationic Ion-Exchange Resin as
Catalyst. React. Funct. Polym. 1993, 20, 1.
(10) Sharma, M. M. Some Novel Aspects of Cationic Ion Exchange
Resins as Catalysts. React. Funct. Polym. 1995, 26, 1.

(11) Harmer, M. A.; Sun, Q. Solid Acid Catalysis Using Ion Exchange
Resins. Appl. Catal., A 2001, 221, 45.
(12) Saigo, K.; Hashimoto, Y.; Hayashi, M. Catalysts for Bisalkoxycarbonylation of Olefins and Manufacture of Succinic Acid Esters. Jpn.
Kokai Tokkyo Koho 2000, 200271485.
(13) Benedict, D. J.; Parulekar, S. J.; Tsai, S. P. Pervaporation-Assisted
Esterification of Lactic and Succinic Acids with Downstream Recovery. J.
Membr. Sci. 2006, 281, 435.
(14) Gangadwala, J.; Mankar, S.; Mahajani, S. Esterification of Acetic
Acid with Butanol in the Presence of Ion-Exchange Resins as Catalysts.
Ind. Eng. Chem. Res. 2003, 42, 2146.
(15) Asthana, N.; Kolah, A.; Vu, D.; Lira, C. T.; Miller, D. J. A
Continuous Reactive Separation Process for Ethyl Lactate Formation. Org.
Process Res. DeV. 2005, 9, 599.
(16) Kolah, A.; Asthana, N. S.; Vu, D. T.; Lira, C. T.; Miller, D. J.
Reaction Kinetics of the Catalytic Esterification of Citric Acid with Ethanol.
Ind. Eng. Chem. Res. 2007, 46, 3180.
(17) Asthana, N. S.; Kolah, A. K.; Vu, D. T.; Lira, C. T.; Miller, D. J.
A Kinetic Model for the Esterification of Lactic Acid and Its Oligomers.
Ind. Eng. Chem. Res. 2006, 45, 5251.

ReceiVed for reView May 9, 2007


ReVised manuscript receiVed March 31, 2008
Accepted May 6, 2008
IE0706616

Das könnte Ihnen auch gefallen