Вы находитесь на странице: 1из 7

Ind. Eng. Chem. Res.

2006, 45, 5251-5257

5251

A Kinetic Model for the Esterification of Lactic Acid and Its Oligomers
Navinchandra S. Asthana, Aspi K. Kolah, Dung T. Vu, Carl T. Lira, and Dennis J. Miller*
Department of Chemical Engineering and Material Science, Michigan State UniVersity,
East Lansing, Michigan 48824

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

The kinetics of esterifying lactic acid and its oligomers with ethanol over Amberlyst 15 cation-exchange
resin have been determined via batch reaction studies. The kinetic model for esterification of lactic acid and
its oligomers uses nth-order, reversible rate expressions for esterification and oligomerization reactions. Rate
constants, activation energies, and equilibrium constants were generated via regression of experimental results.
Model predictions for esterification of 20%, 50%, and 88% lactic acid solutions in water are in good agreement
with experimental results over a wide range of experimental conditions.
1. Introduction
We recently reported on the esterification of lactic acid with
ethanol to form ethyl lactate in a reactive distillation column.1,2
Reactions were performed using different aqueous solutions of
lactic acid as reactants in the presence of Amberlyst 15 cationexchange resin as a catalyst. Water as a reaction product was
continuously removed from the column in the column distillate
stream, thus driving the reaction toward higher acid conversion
and ester yields. Under optimum reaction conditions, >95% acid
conversion with 65% yield of ethyl lactate was achieved.
At concentrations of >20 wt % in water, lactic acid undergoes
oligomerization reactions to form linear oligomer acids. The
extent of oligomerization is inversely related to the water content
of the solution.3,4 These oligomer acids react with ethanol to
yield oligomer esters. In reactive distillation or any other lactic
acid esterification scheme, oligomer formation and esterification
adversely affect ethyl lactate yield and pose a considerable
challenge in predicting process behavior. For accurate design,
therefore, it is critical to include and characterize oligomer
reactions in the process model.
Many reactive distillation models only consider reaction
equilibrium.5,6 These reaction equilibrium models are useful in
cases where the reaction kinetics are very fast (e.g., where
equilibrium is achieved in a short reaction time) but systematically predict higher conversions than those that are obtained
experimentally in cases where the reaction kinetics are slow.
In those cases, it has been shown that kinetics-based simulations
give more accurate and more reliable predictions.7,8 In lactic
acid esterification, where a series of relatively slow reactions
occurs, a reliable knowledge of the kinetics is, thus, essential
to simulate the process effectively.
There have been several reports describing the kinetics of
lactic acid esterification with different alcohols.9-17 Most of
them involve dilute aqueous solutions of lactic acid (e20 wt
%),8-11 thus obviating the need to account for oligomers in
solution. Others have examined esterification kinetics at high
lactic acid concentrations but have not explicitly included the
role of oligomers. Engin et al.16 studied the esterification of 92
wt % lactic acid solutions with ethanol, accounting for hydrolysis of oligomer acids but ignoring esterification reactions. Zhang
et al.17 made no mention of oligomers in their study of
esterification kinetics of 80 wt % lactic acid. The esterification
kinetics developed by Tanaka et al.14 represent the only prior
* To whom correspondence should be addressed. Tel.: 1-517-3533928. Fax: 1-517-432-1105. E-mail address: millerd@egr.msu.edu.

work that considers higher oligomer acids and their esters.


Although they modeled reaction kinetics and generated rate
constants, they did not clearly describe the catalyst quantity used
in the reaction, and, thus, their results have limited utility.
Furthermore, the limited parametric studies they conducted
hinder the applicability of their rate constants to predict
composition profiles over a wide range of reaction conditions.
Because prior kinetic models do not reliably or completely
predict the behavior of lactic acid oligomers and their esters,
we have conducted and report here a kinetic study describing
rate expressions and rate constants for formation, esterification,
and hydrolysis of oligomer acids and esters over a wide range
of lactic acid concentrations, alcohol contents, temperatures, and
catalyst loadings. The kinetic model presented is useful in both
batch and continuous process designs for lactic acid esterification.
2. Experimental Section
2.1. Chemicals and Catalysts. Lactic acid (20 wt % in
water), ethyl lactate (97%), high-performance liquid chromatography (HPLC)-grade acetonitrile, 0.1 N NaOH solution, 0.1
N KOH solution, and Amberlyst-15 cationic exchange resin
were obtained from Aldrich. A 50 wt % aqueous solution of
lactic acid was procured from Purac. Concentrated lactic acid
(88 wt % in water), HPLC-grade water, and absolute ethanol
were purchased from J.T. Baker. Ethyl lactate was purified to
eliminate water prior to use by treating it with 3 molecular
sieves (dried at 550 C) and then distilling it under vacuum to
obtain 100% purity. The purities of all chemicals were
confirmed via gas chromatography (GC).
Amberlyst 15 cation-exchange resin, which is a copolymer
of divinylbenzene and styrene with 15%-18% cross-linking,
is used as the heterogeneous catalyst in this work. The physical
and chemical properties of the resin have been well-documented.18 The nitrogen Brunauer-Emmett-Teller (BET) surface area, pore volume (<120 nm pore diameters) and average
pore size measurements, as determined in a Micromeritics ASAP
2010 instrument, are 53 m2/g, 0.40 cm3/g, and 30 nm,
respectively. The total cation exchange capacity (CEC) of
Amberlyst 15 is 4.6 mequiv H+ per g dry resin. The resin was
washed and dried at 95 C before its use in reaction.
2.2. Reaction Procedure. The kinetics of esterification
involving 88 and 50 wt % lactic acid solutions with ethanol
were measured in a stirred glass batch reactor (85 cm3). This
reactor was equipped with an outer circulating heating jacket
(to maintain constant temperature inside the reactor) that had

10.1021/ie0513604 CCC: $33.50 2006 American Chemical Society


Published on Web 06/23/2006

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

5252

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006

silicon oil as the circulating fluid. The reactor was also equipped
with a condenser, thermocouple, and sampling port.
At the start of reaction, the reactor was filled with the desired
amount of lactic acid and ethanol and heated to reaction
temperature. A catalyst was added to the reactor through one
of the reactor ports. The point of catalyst addition was
considered as time zero for the reaction. The extent of reaction
was followed by sample withdrawal and analysis at regular
intervals.
All reactions that involved 20 wt % lactic acid, as well as
reactions conducted at temperatures of >80 C for 88 and 50
wt % lactic acid, were conducted in a stirred 300-mL Parr batch
reactor (Model 4561, Parr Instrument Co.). For each experiment,
the reagents and catalyst were charged into the reactor and
heated to the reaction temperature. Stirring commenced after
the temperature was attained; this was noted as zero reaction
time. Samples were withdrawn periodically over the course of
the reaction and analyzed. For all reactions, total concentration
and conversion of lactic acid is reported on a monomer
equivalent basis. The 88 wt % lactic acid solution consists of
monomer acid (L1) and linear oligomer dimer acid (L2), trimer
acid (L3), and tetramer acid (L4); therefore, the concentrations
of L2, L3, and L4 were multiplied by 2, 3, and 4, respectively,
and added to the L1 concentration to give the monomer
equivalent concentration.
2.3. Analysis. All samples were analyzed for water (W) and
ethanol (EtOH) using a Varian model 3700 gas chromatograph
that was equipped with a thermal conductivity detector (TCD)
and a stainless steel column (4 m 3.25 mm) packed with a
liquid stationary phase of Porapak Q. The column oven was
subjected to a temperature program that involved heating from
413 K (after a 2-min hold) to 493 K (and held for 6 min) at a
rate of 20 K/min. High-purity helium (99.999% pure) was used
as the carrier gas at a flow rate of 20 mL/min. The injector and
detectors were maintained at 493 K.
Lactic acid monomer (L1) and its linear oligomer acids (L2,
L3, and L4), ethyl lactate (L1E), and ethyl esters of oligomer
acids (L2E, L3E, and L4E) were analyzed using HPLC. A
complete description of the analysis methodology is presented
in detail elsewhere.1-3 The total acid content was determined
by titration with a standard 0.1 N NaOH solution, using
phenolphthalein as an indicator, as a check of HPLC results.
The results obtained for the acid content from HPLC were
comparable to those from titration.
3. Results
The base case conditions for reaction included 88 wt %
lactic acid feed and ethanol with feed quantities of 0.30 mol of
lactic acid (L1 equivalent), 0.2 mol of water (present in 88 wt
% lactic acid solution), and 0.97 mol of ethanol, a catalyst
loading of 3 wt % of total reactant weight, a temperature of 80
C, and an agitation speed of 740 rpm. To cover a broad range
of conditions, esterification reactions were also performed on
solutions containing 20, 50, or 88 wt % lactic acid (compositions
provided in Table 1) using ethanol:lactic acid molar feed ratios
from 1:1 to 4:1, catalyst loadings from 0 wt % to 5 wt %, and
reaction temperatures from 335 K to 363 K. The results of these
experiments were used in the regression analysis of the kinetic
model (described below) to determine an optimum set of rate
constants that describe esterification. Representative results from
experiments and from the kinetic model are represented in
Figures 1-6 by data points and dotted lines, respectively.
Because the 88 wt % lactic acid solution contains the highest
concentrations of oligomers (Table 1) and, thus, represents the

Figure 1. Effect of the ethanol:lactic acid initial molar ratio (MR) on


conversion. Reaction conditions: 88 wt % lactic acid solution, catalyst
loading ) 3 wt %, agitation rate ) 740 rpm, and reaction temperature )
80 C. Dotted lines are model fit; symbols represent experimental data (([)
MR ) 1, (2) MR ) 2, (9) MR ) 3, and (b) MR ) 4).
Table 1. Composition of Lactic Acid Feedstocks
Feed Designation (Nominal)
parameter
feed component
L1

20 wt %
lactic acid

50 wt %
lactic acid

88 wt %
lactic acid

23 wt %
(5.6 mol %)

46 wt %
(15.2 mol %)
3 wt %
(0.5 mol %)

77 wt %
(94.4 mol %)
2.6 M

51 wt %
(84.3 mol %)
5.9 M

58 wt %
(43.5 mol %)
22 wt %
(9.2 mol %)
6 wt % (1.8)
2 wt % (0.4)
12 wt %
(45.1 mol %)
10.8 M

L2
L3
L4
H2O
monomer equivalent
concentration

most-complex reaction mixture, the results from its esterification


are presented in detail here.
3.1. Effect of Reactant Molar Ratio. The effect of the
ethanol:lactic molar feed ratio was studied over the range of
1:1 to 4:1. An increase in lactic acid conversion was observed
with increasing molar ratios (Figure 1) up to 3:1. Little
enhancement in acid conversion was observed when the feed
molar ratio was increased from 3:1 to 4:1; therefore, most
reactions were conducted at a feed molar ratio of 3:1.
3.2. Effect of Catalyst Loading. The effect of resin catalyst
loading on the esterification rate over the range of 0-5 wt %
of total solution mass was studied. The absolute initial rate of
acid conversion was observed to increase linearly with catalyst
loading (Figure 2). From this result, a turnover frequency (TOF)
of 0.06 s-1 was determined for Amberlyst 15 resin at 80 C.
Because lactic acid esterification is autocatalytic, a nonzero
intercept was observed in Figure 2. For the 88 wt % lactic acid
solution, this uncatalyzed esterification rate contributed 20%
of the rate observed for 1% catalyst loading, a contribution that
is equivalent to 0.2 wt % resin catalyst. This contribution is
smaller for lower acid concentrations and is relatively unimportant for higher catalyst loadings; thus, it is omitted from the
kinetic model described below. The effect of catalyst loading

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006 5253

Figure 2. Initial rate of acid conversion versus catalyst loading. Reaction


conditions: 88 wt % lactic acid solution, ethanol:lactic acid molar ratio )
3, agitation rate ) 740 rpm, and reaction temperature ) 80 C.
Figure 4. Esterification of 88 wt % lactic acid solution. Reaction
conditions: ethanol:lactic acid molar ratio ) 3, catalyst loading ) 3 wt %,
agitation rate ) 740 rpm, and reaction temperature ) 80 C. Legend: ([)
L1, (2) L1E, (+) water, (*) ethanol, (9) L2, () L2E, (b) L3, and (0) L3E.

Figure 3. Arrhenius plot of rate constants from initial rate of lactic acid
esterification. Reaction conditions: 88 wt % lactic acid solution, ethanol:
lactic acid molar ratio ) 3, catalyst loading ) 3 wt %, and agitation rate
) 740 rpm. Legend: ([) k1, (9) k2, and (2) k3.

was also observed on reactions of oligomers such as L2E: when


loading was increased from 0 wt % to 5 wt %, the rates of
formation and hydrolysis of L2E were observed to increase
proportionally. Most reactions were performed with a catalyst
loading of 3 wt % or 0.03 kg catalyst/kg solution.
3.3. Effect of Reaction Temperature. The influence of the
reaction temperature on the esterification rate was studied over
the range of 62-90 C. An Arrhenius plot for the initial rate of
reactions 1-3 is given in Figure 3. The activation energy for

Figure 5. Esterification of 50 wt % lactic acid solution. Reaction


conditions: ethanol:lactic acid molar ratio ) 3, catalyst loading ) 3 wt %,
agitation rate ) 740 rpm, and reaction temperature ) 80 C. Legend: ([)
L1, (2) L1E, (9) L2, () L2E, (*) water, and (b) ethanol.

the esterification reactions (reactions 1-3), based on the initial


rate, increased with the degree of oligomerization of lactic

5254

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006

pletely described by reactions 1, 2, and 4 and by reaction 1,


respectively.
k1

L1 + EtOH y\
z L1E + H2O
k /K
1

k2

L2 + EtOH y\
z L2E + H2O
k /K
2

k3

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

k4

k5

4. Kinetic Model
4.1. Reaction Pathways. Based on the reactant and product
composition profiles during esterification of the 88 wt % lactic
acid solution, a set of reaction pathways have been defined for
the kinetic model. The set of reactions, which are given in
reactions 1-5 below, describes lactic acid monomer esterification as well as oligomer formation and esterification. All
oligomers that are larger than dimers (L2 and L2E) are lumped
together as L3 and L3E; the quantities of larger oligomers are
small enough that they do not significantly influence the
reaction, either kinetically or with regard to overall acid
concentration and mass balances. A similar set of reaction
pathways was proposed by Tanaka et al.14 Esterification of
the 50 wt % and 20 wt % lactic acid solutions is com-

(4)

L3 + H2O y\
z L 1 + L2
k /K

3.4. Effect of Lactic Acid Feed Composition. Component


profiles for esterification of 88, 50, and 20 wt % lactic acid
feed solutions, under typical conditions, are given in Figures
4-6. Because of the presence of oligomers, it is apparent that
predicting the behavior of the 88 wt % lactic acid solution
requires a more-complex model than that required for 50 wt %
and 20 wt % feed solutions.

(3)

z 2L1
L2 + H2O y\
k /K

acids: from 48 000 kJ/kmol for esterification of monomer lactic


acid (L1) to 74 100 kJ/kmol for esterification of trimer lactic
acid (L3). These values are consistent with kinetically controlled
reactions. Further reactions were performed at 80 C, which is
the normal boiling point of ethanol.

(2)

L3 + EtOH y\
z L3E + H2O
k /K

Figure 6. Esterification of 20 wt % lactic acid solution. Reaction


conditions: ethanol:lactic acid molar ratio ) 3, catalyst loading ) 3 wt %,
agitation rate ) 740 rpm, and reaction temperature ) 80 C. Legend: ([)
L1, (2) L1E, (9) ethanol, and () water.

(1)

(5)

4.2. Evaluation of Mass-Transfer Limitations. Two possible modes of mass-transfer limitations are associated with
heterogeneously catalyzed reactions: one across the solid/liquid
interface (external mass transfer) and other within the catalyst
particle (intraparticle diffusion). In the evaluations below, liquidphase effective diffusivities were calculated using the WilkeChang equation.19
Chakrabarti and Sharma20 suggested that external masstransfer limitations for ion-exchange resin-catalyzed reactions
can be eliminated by conducting experiments at stirring rates
of >500 rpm. To check their assumption, a theoretical calculation was performed to ascertain the magnitude of the external
mass-transfer rate. Solid-liquid mass-transfer coefficients were
calculated using a Sherwood number (kSL,jLt/De,j) of 2.0, which
is the theoretical minimum value. The observed reaction rates
of L1, L2, and L3 + L4 were calculated from their concentrationversus-time profiles. (Oligomers L3 and L4 were lumped together
because of uncertainty in the response factor for L4 and because
of its low concentration.) For each component, the maximum
mass-transfer rate (kSL,j Cj) was determined to be at least an
order of magnitude greater than the observed reaction rate,
indicating that there are no significant solid-liquid mass-transfer
limitations. Resin particles physically disintegrate at stirring rates
of >800 rpm; therefore, all reactions were conducted at a speed
of 740 rpm.
The possible presence of intraparticle mass-transfer resistances
was examined by determining the observable modulus (2 )
(-rj)Lt2/Cj/De,j) and applying the Weisz-Prater criterion; the
maximum value of 2 observed was 0.15, which is sufficiently
small to indicate that intraparticle mass-transport resistances are
unimportant. This is consistent with previous reports that
intraparticle diffusion limitations in reactions with Amberlyst15 under similar conditions could be considered negligible.9,10,21
4.3. Kinetic Model Equations. A simple nth-order reversible
reaction model has been chosen to describe lactic acid esterification that has been catalyzed by Amberlyst 15 ion-exchange
resin. The rate of formation for each component in the reaction

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006 5255

system described by reactions 1-5 above is given in eqs 6-13


below.

dxL1

dt

[ ( )(
[ ( )(
EA1

) wcat k10 exp -

xL1xEtOH -

RT

EA4 xL12

wcat k4 exp 0

RT

K4

K1

EA5 xL1xL2

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

dxL2
dt

RT

[ ( )(
[ ( )(
EA2

) wcat k20 exp -

E A4

wcat k4 exp 0

RT

K5

K2

xL2xW -

( )(

xL12
K4

EA5 xL1xL2

k50 exp -

dxL3
dt

RT

[ ( )(
EA3

) wcat k30 exp -

RT

xL3xEtOH E A5

E A1

) wcat k10 exp -

dt

dt
dxL3E
dt

K5

[ (
[ (
[ (
[ (

RT

)(
)(
)(

E A2

) wcat k20 exp -

RT

E A3

) wcat k30 exp -

RT

xL2xEtOH xL3xEtOH -

(6)

- xL3xW

)]

xL3xW -

xL1xEtOH -

)]

K3

[ ( )(

dxL1E

dxL2E

RT

)]

xL3ExW

[ ( )(

wcat k50 exp -

- xL3xW

xL2ExW

xL2xEtOH -

RT

- xL2xW +

( )(

k50 exp -

)]

xL1ExW

)]

)]

xL1xL2
K5

(8)

)]

xL1ExW
K1

(9)

)]
)]
)]

xL2ExW
K2

(10)

xL3ExW
K3

(7)

[ ( )(
[ ( )(

EA1 xL1ExW
dxW
- xL1xEtOH
) wcat k10 exp dt
RT
K1
EA2 xL2ExW

wcat k20 exp -

( )(

EA3 xL3ExW

k30 exp -

RT

xL2xW -

K3

xL12
K4

RT

K2

)]
)

)]

- xL2xEtOH +

)] [ ( )
( )(
)]
E A4

- xL3xEtOH

+ wcat k40 exp -

E A5

+ k50 exp -

RT

xL3xW +

xL1xL2
K5

parameter

RT

(13)

value

k01
k02
k03
k04
k05

1.91 105 kgsol/kgcat/s


2.66 104 kgsol/kgcat/s
1.24 107 kgsol/kgcat/s
1.62 104 kgsol/kgcat/s
6.67 103 kgsol/kgcat/s

EA1
EA2
EA3
EA4
EA5

48000 kJ/kmol
54500 kJ/kmol
74100 kJ/kmol
52000 kJ/kmol
50800 kJ/kmol

K1
K2
K3
K4
K5

2.4
0.6
0.3
5.0
5.0

The equilibrium constant for L1E formation (K1) was obtained


via the reaction of 20 wt % solution of lactic acid with ethanol
in the presence of Amberlyst 15. Complete chemical equilibrium
was achieved after 72 h of reaction. The value for K1 was
determined to be 2.4. The equilibrium constants for hydrolysis
of all linear acid oligomers were taken as 5.0, based on our
earlier work.3 The equilibrium constants for L2E and L3E
formation via esterification of L2 and L3 were treated as
adjustable parameters in the model.
The pre-exponential factors (k0i ) and activation energies
(EAi) in the kinetic model were fit to the experimental data. To
minimize the number of degrees of freedom, the activation
energies for reactions 1-3 were fixed at the values obtained
from Figure 2 for initial reaction rates. Values for the five preexponential factors and EA4 and EA5 were then generated by
nonlinear regression analysis using MATLAB 7.0 to minimize
the sum of the mean-square differences (eq 14) between
experimental and predicted liquid-phase mole fractions for all
species over the course of reaction. The resulting pre-exponential
factors, activation energies, and equilibrium constants for all
reactions are given in Table 2.

Fmin2 )

(11)

E A1
xL1ExW
dxEtOH
) wcat k10 exp xL1xEtOH +
dt
RT
K1
E A2
xL2ExW
xL2xEtOH +
wcat k20 exp RT
K2
E A3
xL3ExW
k30 exp xL3xEtOH (12)
RT
K3

)(
( )(

Table 2. Parameters of the Kinetic Model

(xi,cal - xi,expt)2

samples
nsamples

(14)

5. Discussion
To illustrate the validity of the kinetic model, composition
profiles were generated from eqs 6-13 for the entire range of
experimental conditions examined. Representative results of the
model are shown in Figures 1-6. (Additional esterification
results are given in the Supporting Information.) For 50 and 88
wt % lactic acid, predictions for component profiles were in
reasonably good agreement with the experimental values; for
20 wt % lactic acid, predictions were in excellent agreement
with experiments. Error values or deviations (expressed as a
percentage) between experimental and model predictions for
each component are provided in Table 3. For experiments with
88 wt % feed solution, errors were small during the initial 180
min of reaction but became larger as the reaction approached
equilibrium. Errors were larger for L2E and L3E on a percentage
basis, but those species were present in small quantities and,
thus, do not significantly affect major component concentrations.
For 88 wt % lactic acid, overprediction of L2E formation at
equilibrium resulted in overprediction of acid conversion relative
to the experiment (Figure 3).
The kinetic model described here is based on the concentration of species (e.g., mole fraction) in solution. We recognize

5256

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006

Table 3. Comparison of Experimental Values and Model Prediction for Esterification of 20, 50, and 88 wt % Lactic Acid Aqueous Solutions
with Ethanol

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

EtOH:LA
molar ratioa

catalyst
loading (wt %)

reaction
temperature (C)

L1

L2

L3

Percent Error
L1E
L2E

L3E

EtOH

20.5
10.1
7.7
2.7
11.2
7
3.6
2.6
2.5
0.6
6.6

37.1
15.9
12.2
16.9
16
9
5.7
6.6
9.6
3.6
14.5

25.1
28
20.5
25
31.2
29.8
29.1
16
31.5
13.1
29.6

10.9
4.4
2
1.1
1.6
1.4
1.4
1.2
0.8
0.6
3

5.2
2.3
2.1
0.9
2.1
1.6
1
0.6
0.8
0.2
4.2

49
29.2
37.1
25.9
14.7
24.4
24.7
39.1
14.8
34.8

6.2
2.9
1.1
0.5
2.0
1
1.4
0.4
1.2
1.2

0.4
0.04
0.4
0.6
0.2
0.4
0.4
0.8
0.6
0.5

0.8
0.3
0.2
0.1
1.3
0.2
0.7
0.2
0.5
0.2

0.06
0.05
0.07
0.2
0.4
0.03
0.2
0.08
0.2
0.05

1
2
3
4
3
3
3
3
3
3
3

3
3
3
3
1
2
4
5
3
3
3

80
80
80
80
80
80
80
80
62
72
90

88 wt % Lactic Acid Solution


8.7
19.9
24
8.9
17.2
15.5
13.9
15.8
11.2
22.2
21
12.8
11.3
9.8
9.3
7.4
10.1
7.6
15.7
7
4.8
11.5
33.4
27.9
8.3
1.2
1.8
5.5
3.5
3.2
9.5
41.8
25.9

1
2
3
4
3
3
3
3
3
3

3
3
3
3
1
2
4
3
3
3

80
80
80
80
80
80
80
61
71
90

50 wt % Lactic Acid Solution


2.6
6.9
2.2
5.1
5.1
10.3
8.4
20.4
4.2
1.8
5.6
6
6.3
14
6.6
3.2
7.1
8.4
1
20.8

20.2
11.2
4.8
5
13
4
3.5
14.2
5.2
4.1

1
2
3
4
3
3
3
3
3
3

3
3
3
3
1
2
4
3
3
3

80
80
80
80
80
80
80
60
70
90

20 wt % Lactic Acid Solution


1.5
1.4
1.8
4.4
5.8
1
2
1.2
2.5
1.2

9.2
3.5
4.2
17
9.7
5.8
14.1
2.8
5.6
4.6

Molar ratio of ethanol to lactic acid.

that species activity is a fundamentally more appropriate variable


for kinetic modeling; however, for this system, there is little or
no experimental data available on liquid-phase activity coefficients of lactic acid, its oligomers, and their esters in ethanolwater solutions. Prediction of such activity coefficients, which
is possible using UNIFAC or other group contribution methods,
is subject to substantial uncertainties, particularly for the high
concentrations that have been examined here. In fact, we did
prepare an activity-based kinetic model for the lactate esterification system that included the application of UNIFAC for
activity coefficient prediction. The optimized activity-based
model gave errors larger than those obtained with the mole
fraction-based model; thus, we do not believe there is any
advantage to application of such a model at this time.
The choice of simple nth-order kinetic expressions for
esterification, as opposed to a more mechanistically based
approach (such as a Langmuir-Hinshelwood model), was made
partially because of uncertainty in the liquid-phase environment
within the Amberlyst 15 cation-exchange resin. When placed
into aqueous solutions that contain ethanol and organic acid,
Amberlyst 15 swells, because of the selective absorption of
water.22 We observe the same behavior with dilute solutions of
lactic acid (20 wt % and 50 wt %) in esterification, where water
is present in large excess, relative to lactic acid (20-fold excess
for 20 wt % and 6-fold excess for 50 wt %), and, thus, free
water can absorb. When concentrated (88 wt %) lactic acid,
where the molar ratio of water to lactic acid is 0.7, is subject to
esterification, ethanol rather than water is preferentially absorbed
into Amberlyst 15. We attribute this behavior to the association
of free water with lactic acid, such that very little or no free

water is available for absorption onto resin particles at high lactic


acid concentration. This hypothesis was further verified by an
experiment in which a known composition of 88 wt % lactic
acid and ethanol was brought in contact with the resin at room
temperature, equilibrated for 30 min, and analyzed. At this
temperature, no reaction was observed and ethanol was observed
to be preferentially adsorbed.
6. Conclusions
The kinetics of esterification of lactic acid and its linear
oligomers with ethanol have been described using simple nthorder reversible rate expressions. Pre-exponential factors, activation energies, and equilibrium constants were either fit to the
experimental data using nonlinear regression analysis or, in some
cases, were calculated directly or taken from other work.
Activation energies in the range of 48 000-74 100 kJ/kmol and
analysis of mass-transport resistances indicate that the reaction
is kinetically controlled. The predicted time-dependent species
profiles in esterification were in good agreement with experimental results over a wide range of reaction conditions.
Acknowledgment
Financial support from National Corn Growers Association
and the U.S. Department of Energy are greatly appreciated. The
authors also appreciate assistance from MSCRTSF at Michigan
State University for mass spectroscopic analysis and identification of various components.
Supporting Information Available: Figures containing
experimental data and model fit for additional lactic acid

Ind. Eng. Chem. Res., Vol. 45, No. 15, 2006 5257

esterification studies (PDF, 28 pages). This material is available


free of charge via the Internet at http://pubs.acs.org.

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


Publication Date (Web): June 23, 2006 | doi: 10.1021/ie0513604

Nomenclature and Units


Cj ) concentration of species j in liquid-phase solution (gmol/kgsoln)
De,j ) effective diffusivity of species j in resin (m2/s)
EAi ) energy of activation for the ith reaction (kJ/kmol)
ki0 ) pre-exponential factor for the ith reaction (kgsol/kgcat/s)
kSL,j ) mass-transfer coefficient of species j (m/s)
Ki ) equilibrium constant for the ith reaction
Lt ) resin particle size equivalent; Lt ) D/6 ) 1 10-4 m
R ) gas constant; R ) 8.31 kJ/(kmol K)
rj ) initial reaction rate of species j (g-mol/kgsoln/s)
T ) temperature (K)
wcat ) catalyst concentration (kgcat/kgsoln)
xj ) mole fraction of species j in liquid-phase solution
Greek Letters
2 ) observable modulus
Literature Cited
(1) Asthana, N.; Kolah, A.; Vu, D.; Lira, C. T.; Miller, D. J. A
Continuous Reactive Separation Process for Ethyl Lactate Formation. Org.
Proc. Res. DeV. 2005, 9, 599.
(2) Asthana, N.; Kolah, A.; Lira, C. T.; Miller, D. J. Improved Process
for Production of Organic Acid Esters. U.S. Patent Application No. 2006/
0014977 A1, 2004.
(3) Vu, D. T.; Kolah, A. K.; Asthana, N. S.; Peereboom, L.; Lira, C.
T.; Miller, D. J. Oligomer Distribution in Concentrated Lactic acid Solutions.
Fluid Phase Equilib. 2005, 236, 39.
(4) Holten, C. H. Lactic Acid: Properties and Chemistry of Lactic Acid
and DeriVatiVes; Verlag Chemie: Copenhagen, 1971.
(5) Taylor, R.; Krishna, R. Modelling Reactive Distillation. Chem. Eng.
Sci. 2000, 55, 5183.
(6) Popken, T.; Gotze, L.; Gmehling, J. Reaction Kinetics and Chemical
Equilibrium of Homogeneously and Heterogeneously Catalyzed Acetic Acid
Esterification with Methanol and Methyl Acetate Hydrolysis. Ind. Eng.
Chem. Res. 2000, 39, 2601.
(7) Dhale, A. D.; Myrant, L. K.; Chopade, S. P.; Jackson, J. E.; Miller,
D. J. Propylene Glycol and Ethylene Glycol Recovery from Aqueous
Solutions via Reactive Distillation. Chem. Eng. Sci. 2004, 59, 2881.

(8) Beckmann, A.; Nierlich, F.; Popeken, T.; Reusch, D.; von Scala,
C.; Tuchlenski, A. Industrial Experience in Scale-up of Reactive Distillation
with Examples from C4-Chemistry. Chem. Eng. Sci. 2002, 57, 1525.
(9) Sanz, M. T.; Murga, R.; Beltran, S.; Cabezas, J. L.; Coca, J.
Autocatalyzed and Ion-Exchanged-Resin-Catalyzed Esterification Kinetics
of Lactic acid with Methanol. Ind. Eng. Chem. Res. 2002, 41, 512.
(10) Sanz, M. T.; Murga, R.; Beltran, S.; Cabezas, J. L.; Coca, J. Kinetic
Study for Reactive System of Lactic Acid Esterification with Methanol:
Methyl Lactate Hydrolysis Reaction. Ind. Eng. Chem. Res. 2004, 43, 2049.
(11) Choi, J. I.; Hong, W. H.; Chang, H. N. Reaction Kinetics of Lactic
Acid with Methanol Catalyzed by Acid Resins. Int. J. Chem. Kinet. 1996,
28, 37.
(12) See, Y.; Hong, W. H.; Hong, T. H. Effects of Operation Variables
on the Recovery of Lactic Acid in a Batch Distillation Process with Chemical
Reactions. Korean J. Chem. Eng. 1999, 16, 556.
(13) Yadav, G. D.; Kulkarni, H. B. Ion-Exchange Resin Catalysis in
Synthesis of Isopropyl Lactate. React. Funct. Polym. 2000, 44, 153.
(14) Tanaka, K.; Yoshikawa, R.; Ying, C.; Kita, H.; Okamoto, K.
Application of Zeolite T Membrane to Vapor-Permeation-Aided Esterification of Lactic Acid with Ethanol. Chem. Eng. Sci. 2002, 57, 1577.
(15) Benedict, D. J.; Parulekar, S. J.; Tsai, S. Esterification of Lactic
Acid and Ethanol with/without Pervaporation. Ind. Eng. Chem. Res. 2003,
42, 2282.
(16) Ayturk, E.; Hamachi, H.; Karakas, G. Production of Lactic Acid
Esters Catalyzed by Heteropolyacid Supported Over Ion-Exchange Resins.
Green Chem. 2003, 5, 460.
(17) Zhang, Y.; Ma, L.; Yang, J. Kinetics of Esterification of Lactic
Acid with Ethanol Catalyzed by Cation-Exchange Resin. React. Funct.
Polym. 2004, 61, 101.
(18) Rohm & Hass. Product Data Sheet, Amberlyst-15 Dry. Available
via the Internet at http://www.rohmhaas.com/ionexchange/IP/sac.htm.
(19) Reid, R. C.; Prausnitz, M. J.; Sherwood, T. K. The Properties of
Gases and Liquids; McGraw-Hill: New York, 1987.
(20) Chakrabarti, A.; Sharma, M. M. Cationic Ion-Exchange Resin as
Catalyst. React. Polym. 1993, 20, 1.
(21) Gangadwala, J.; Mankar, S.; Mahajani, S.; Kienle, A.; Stein, E.
Esterification of Acetic Acid with Butanol in the Presence of Ion-Exchange
Resins as Catalysts. Ind. Eng. Chem. Res. 2003, 42, 2146.
(22) Mazzotti, M.; Neri, B.; Gelosa, D.; Kruglov, A.; Morbidelli, M.
Kinetics of Liquid-Phase Esterification Catalyzed by Acidic Resins. Ind.
Eng. Chem. Res. 1997, 36, 3.

ReceiVed for reView December 6, 2005


ReVised manuscript receiVed April 18, 2006
Accepted May 10, 2006
IE0513604

Вам также может понравиться