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

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 work that considers higher oligomer acids and their esters.
Although they modeled reaction kinetics and generated rate
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

We recently reported on the esterification of lactic acid with constants, they did not clearly describe the catalyst quantity used
Downloaded via UNIV NACIONAL DE COLOMBIA on April 21, 2019 at 01:22:45 (UTC).

ethanol to form ethyl lactate in a reactive distillation column.1,2 in the reaction, and, thus, their results have limited utility.
Reactions were performed using different aqueous solutions of Furthermore, the limited parametric studies they conducted
lactic acid as reactants in the presence of Amberlyst 15 cation- hinder the applicability of their rate constants to predict
exchange resin as a catalyst. Water as a reaction product was composition profiles over a wide range of reaction conditions.
continuously removed from the column in the column distillate Because prior kinetic models do not reliably or completely
stream, thus driving the reaction toward higher acid conversion predict the behavior of lactic acid oligomers and their esters,
and ester yields. Under optimum reaction conditions, >95% acid we have conducted and report here a kinetic study describing
conversion with 65% yield of ethyl lactate was achieved. rate expressions and rate constants for formation, esterification,
At concentrations of >20 wt % in water, lactic acid undergoes and hydrolysis of oligomer acids and esters over a wide range
oligomerization reactions to form linear oligomer acids. The of lactic acid concentrations, alcohol contents, temperatures, and
extent of oligomerization is inversely related to the water content catalyst loadings. The kinetic model presented is useful in both
of the solution.3,4 These oligomer acids react with ethanol to batch and continuous process designs for lactic acid esterifica-
yield oligomer esters. In reactive distillation or any other lactic tion.
acid esterification scheme, oligomer formation and esterification
adversely affect ethyl lactate yield and pose a considerable
2. Experimental Section
challenge in predicting process behavior. For accurate design,
therefore, it is critical to include and characterize oligomer 2.1. Chemicals and Catalysts. Lactic acid (20 wt % in
reactions in the process model. water), ethyl lactate (97%), high-performance liquid chroma-
Many reactive distillation models only consider reaction tography (HPLC)-grade acetonitrile, 0.1 N NaOH solution, 0.1
equilibrium.5,6 These reaction equilibrium models are useful in N KOH solution, and Amberlyst-15 cationic exchange resin
cases where the reaction kinetics are very fast (e.g., where were obtained from Aldrich. A 50 wt % aqueous solution of
equilibrium is achieved in a short reaction time) but systemati- lactic acid was procured from Purac. Concentrated lactic acid
cally predict higher conversions than those that are obtained (88 wt % in water), HPLC-grade water, and absolute ethanol
experimentally in cases where the reaction kinetics are slow. were purchased from J.T. Baker. Ethyl lactate was purified to
In those cases, it has been shown that kinetics-based simulations eliminate water prior to use by treating it with 3 Å molecular
give more accurate and more reliable predictions.7,8 In lactic sieves (dried at 550 °C) and then distilling it under vacuum to
acid esterification, where a series of relatively slow reactions obtain ∼100% purity. The purities of all chemicals were
occurs, a reliable knowledge of the kinetics is, thus, essential confirmed via gas chromatography (GC).
to simulate the process effectively. Amberlyst 15 cation-exchange resin, which is a copolymer
There have been several reports describing the kinetics of of divinylbenzene and styrene with 15%-18% cross-linking,
lactic acid esterification with different alcohols.9-17 Most of is used as the heterogeneous catalyst in this work. The physical
them involve dilute aqueous solutions of lactic acid (e20 wt and chemical properties of the resin have been well-docu-
%),8-11 thus obviating the need to account for oligomers in mented.18 The nitrogen Brunauer-Emmett-Teller (BET) sur-
solution. Others have examined esterification kinetics at high face area, pore volume (<120 nm pore diameters) and average
lactic acid concentrations but have not explicitly included the pore size measurements, as determined in a Micromeritics ASAP
role of oligomers. Engin et al.16 studied the esterification of 92 2010 instrument, are 53 m2/g, 0.40 cm3/g, and 30 nm,
wt % lactic acid solutions with ethanol, accounting for hydroly- respectively. The total cation exchange capacity (CEC) of
sis of oligomer acids but ignoring esterification reactions. Zhang Amberlyst 15 is 4.6 mequiv H+ per g dry resin. The resin was
et al.17 made no mention of oligomers in their study of washed and dried at 95 °C before its use in reaction.
esterification kinetics of 80 wt % lactic acid. The esterification 2.2. Reaction Procedure. The kinetics of esterification
kinetics developed by Tanaka et al.14 represent the only prior involving 88 and 50 wt % lactic acid solutions with ethanol
were measured in a stirred glass batch reactor (85 cm3). This
* To whom correspondence should be addressed. Tel.: 1-517-353- reactor was equipped with an outer circulating heating jacket
3928. Fax: 1-517-432-1105. E-mail address: millerd@egr.msu.edu. (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
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 Figure 1. Effect of the ethanol:lactic acid initial molar ratio (MR) on
of L2, L3, and L4 were multiplied by 2, 3, and 4, respectively, conversion. Reaction conditions: 88 wt % lactic acid solution, catalyst
and added to the L1 concentration to give the monomer loading ) 3 wt %, agitation rate ) 740 rpm, and reaction temperature )
80 °C. Dotted lines are model fit; symbols represent experimental data (([)
equivalent concentration.
MR ) 1, (2) MR ) 2, (9) MR ) 3, and (b) MR ) 4).
2.3. Analysis. All samples were analyzed for water (W) and
ethanol (EtOH) using a Varian model 3700 gas chromatograph Table 1. Composition of Lactic Acid Feedstocks
that was equipped with a thermal conductivity detector (TCD) Feed Designation (Nominal)
and a stainless steel column (4 m × 3.25 mm) packed with a 20 wt % 50 wt % 88 wt %
liquid stationary phase of Porapak Q. The column oven was parameter lactic acid lactic acid lactic acid
subjected to a temperature program that involved heating from feed component
413 K (after a 2-min hold) to 493 K (and held for 6 min) at a L1 23 wt % 46 wt % 58 wt %
rate of 20 K/min. High-purity helium (99.999% pure) was used (5.6 mol %) (15.2 mol %) (43.5 mol %)
L2 3 wt % 22 wt %
as the carrier gas at a flow rate of 20 mL/min. The injector and (0.5 mol %) (9.2 mol %)
detectors were maintained at 493 K. L3 6 wt % (1.8)
Lactic acid monomer (L1) and its linear oligomer acids (L2, L4 2 wt % (0.4)
H2O 77 wt % 51 wt % 12 wt %
L3, and L4), ethyl lactate (L1E), and ethyl esters of oligomer (94.4 mol %) (84.3 mol %) (45.1 mol %)
acids (L2E, L3E, and L4E) were analyzed using HPLC. A monomer equivalent 2.6 M 5.9 M 10.8 M
concentration
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 most-complex reaction mixture, the results from its esterification
phenolphthalein as an indicator, as a check of HPLC results. are presented in detail here.
The results obtained for the acid content from HPLC were 3.1. Effect of Reactant Molar Ratio. The effect of the
comparable to those from titration. ethanol:lactic molar feed ratio was studied over the range of
1:1 to 4:1. An increase in lactic acid conversion was observed
3. Results with increasing molar ratios (Figure 1) up to 3:1. Little
The “base case” conditions for reaction included 88 wt % enhancement in acid conversion was observed when the feed
lactic acid feed and ethanol with feed quantities of 0.30 mol of molar ratio was increased from 3:1 to 4:1; therefore, most
lactic acid (L1 equivalent), 0.2 mol of water (present in 88 wt reactions were conducted at a feed molar ratio of ∼3:1.
% lactic acid solution), and 0.97 mol of ethanol, a catalyst 3.2. Effect of Catalyst Loading. The effect of resin catalyst
loading of 3 wt % of total reactant weight, a temperature of 80 loading on the esterification rate over the range of 0-5 wt %
°C, and an agitation speed of 740 rpm. To cover a broad range of total solution mass was studied. The absolute initial rate of
of conditions, esterification reactions were also performed on acid conversion was observed to increase linearly with catalyst
solutions containing 20, 50, or 88 wt % lactic acid (compositions
loading (Figure 2). From this result, a turnover frequency (TOF)
provided in Table 1) using ethanol:lactic acid molar feed ratios
of 0.06 s-1 was determined for Amberlyst 15 resin at 80 °C.
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 Because lactic acid esterification is autocatalytic, a nonzero
experiments were used in the regression analysis of the kinetic intercept was observed in Figure 2. For the 88 wt % lactic acid
model (described below) to determine an optimum set of rate solution, this uncatalyzed esterification rate contributed ∼20%
constants that describe esterification. Representative results from of the rate observed for 1% catalyst loading, a contribution that
experiments and from the kinetic model are represented in is equivalent to ∼0.2 wt % resin catalyst. This contribution is
Figures 1-6 by data points and dotted lines, respectively. smaller for lower acid concentrations and is relatively unim-
Because the 88 wt % lactic acid solution contains the highest portant for higher catalyst loadings; thus, it is omitted from the
concentrations of oligomers (Table 1) and, thus, represents the kinetic model described below. The effect of catalyst loading
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 Figure 5. Esterification of 50 wt % lactic acid solution. Reaction
loading of 3 wt % or 0.03 kg catalyst/kg solution. conditions: ethanol:lactic acid molar ratio ) 3, catalyst loading ) 3 wt %,
agitation rate ) 740 rpm, and reaction temperature ) 80 °C. Legend: ([)
3.3. Effect of Reaction Temperature. The influence of the
L1, (2) L1E, (9) L2, (×) L2E, (*) water, and (b) ethanol.
reaction temperature on the esterification rate was studied over
the range of 62-90 °C. An Arrhenius plot for the initial rate of the esterification reactions (reactions 1-3), based on the initial
reactions 1-3 is given in Figure 3. The activation energy for 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\
k /K
z L1E + H2O (1)
1 1

k2
L2 + EtOH y\
k /K
z L2E + H2O (2)
2 2

k3
L3 + EtOH y\
k /K
z L3E + H2O (3)
3 3

k4
L2 + H2O y\
k /K
z 2L1 (4)
4 4

k5
L3 + H2O y\
k /K
z L 1 + L2 (5)
5 5

4.2. Evaluation of Mass-Transfer Limitations. Two pos-


sible modes of mass-transfer limitations are associated with
Figure 6. Esterification of 20 wt % lactic acid solution. Reaction heterogeneously catalyzed reactions: one across the solid/liquid
conditions: ethanol:lactic acid molar ratio ) 3, catalyst loading ) 3 wt %, interface (external mass transfer) and other within the catalyst
agitation rate ) 740 rpm, and reaction temperature ) 80 °C. Legend: ([) particle (intraparticle diffusion). In the evaluations below, liquid-
L1, (2) L1E, (9) ethanol, and (×) water.
phase effective diffusivities were calculated using the Wilke-
Chang equation.19

acids: from 48 000 kJ/kmol for esterification of monomer lactic Chakrabarti and Sharma20 suggested that external mass-
acid (L1) to 74 100 kJ/kmol for esterification of trimer lactic transfer limitations for ion-exchange resin-catalyzed reactions
acid (L3). These values are consistent with kinetically controlled can be eliminated by conducting experiments at stirring rates
reactions. Further reactions were performed at 80 °C, which is of >500 rpm. To check their assumption, a theoretical calcula-
the normal boiling point of ethanol. tion was performed to ascertain the magnitude of the external
mass-transfer rate. Solid-liquid mass-transfer coefficients were
3.4. Effect of Lactic Acid Feed Composition. Component calculated using a Sherwood number (kSL,jLt/De,j) of 2.0, which
profiles for esterification of 88, 50, and 20 wt % lactic acid is the theoretical minimum value. The observed reaction rates
feed solutions, under typical conditions, are given in Figures of L1, L2, and L3 + L4 were calculated from their concentration-
4-6. Because of the presence of oligomers, it is apparent that versus-time profiles. (Oligomers L3 and L4 were lumped together
predicting the behavior of the 88 wt % lactic acid solution because of uncertainty in the response factor for L4 and because
requires a more-complex model than that required for 50 wt % of its low concentration.) For each component, the maximum
and 20 wt % feed solutions. 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
4. Kinetic Model limitations. Resin particles physically disintegrate at stirring rates
of >800 rpm; therefore, all reactions were conducted at a speed
of ∼740 rpm.
4.1. Reaction Pathways. Based on the reactant and product
composition profiles during esterification of the 88 wt % lactic The possible presence of intraparticle mass-transfer resistances
acid solution, a set of reaction pathways have been defined for was examined by determining the observable modulus (ηφ2 )
the kinetic model. The set of reactions, which are given in (-rj)Lt2/Cj/De,j) and applying the Weisz-Prater criterion; the
reactions 1-5 below, describes lactic acid monomer esterifi- maximum value of ηφ2 observed was 0.15, which is sufficiently
cation as well as oligomer formation and esterification. All small to indicate that intraparticle mass-transport resistances are
oligomers that are larger than dimers (L2 and L2E) are lumped unimportant. This is consistent with previous reports that
together as L3 and L3E; the quantities of larger oligomers are intraparticle diffusion limitations in reactions with Amberlyst-
small enough that they do not significantly influence the 15 under similar conditions could be considered negligible.9,10,21
reaction, either kinetically or with regard to overall acid 4.3. Kinetic Model Equations. A simple nth-order reversible
concentration and mass balances. A similar set of reaction reaction model has been chosen to describe lactic acid esteri-
pathways was proposed by Tanaka et al.14 Esterification of fication that has been catalyzed by Amberlyst 15 ion-exchange
the 50 wt % and 20 wt % lactic acid solutions is com- 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 Table 2. Parameters of the Kinetic Model
below. parameter value

[ ( )( )]
k01 1.91 × 105 kgsol/kgcat/s
dxL1 EA1 xL1ExW 2.66 × 104 kgsol/kgcat/s
k02
- ) wcat k10 exp - xL1xEtOH - + 1.24 × 107 kgsol/kgcat/s
dt RT K1 k03

[ ( )( )
k04 1.62 × 104 kgsol/kgcat/s
EA4 xL12 k05 6.67 × 103 kgsol/kgcat/s
wcat k4 exp -
0
- xL2xW +
RT K4 EA1 48000 kJ/kmol

( )( )]
EA2 54500 kJ/kmol
EA5 xL1xL2 EA3 74100 kJ/kmol
k50 exp - - xL3xW (6) EA4 52000 kJ/kmol
RT K5 EA5 50800 kJ/kmol

[ ( )( )]
K1 2.4
dxL2 EA2 xL2ExW K2 0.6
- ) wcat k20 exp - xL2xEtOH - + K3 0.3
dt RT K2 K4 5.0

[ ( )( )
K5 5.0
E A4 xL12
The equilibrium constant for L1E formation (K1) was obtained
wcat k4 exp -
0
xL2xW - +
RT K4 via the reaction of 20 wt % solution of lactic acid with ethanol

( )( )]
in the presence of Amberlyst 15. Complete chemical equilibrium
EA5 xL1xL2
was achieved after 72 h of reaction. The value for K1 was
k50 exp - - xL3xW (7)
RT K5 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
dxL3 EA3 xL3ExW
formation via esterification of L2 and L3 were treated as
- ) wcat k30 exp - xL3xEtOH - +
dt RT K3 adjustable parameters in the model.

[ ( )( )]
E A5 xL1xL2 The pre-exponential factors (k0i ) and activation energies
wcat k50 exp - xL3xW - (8) (EAi) in the kinetic model were fit to the experimental data. To
RT K5 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 pre-
dxL1E E A1 xL1ExW
exponential factors and EA4 and EA5 were then generated by
- ) wcat k10 exp - xL1xEtOH - (9)
dt RT K1 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
dxL2E E A2 xL2ExW
reactions are given in Table 2.
- ) wcat k20 exp - xL2xEtOH - (10)
dt RT K2
∑ (xi,cal - xi,expt)2

[ ( )( )]
samples
dxL3E E A3 xL3ExW Fmin2 ) (14)
- ) wcat k30 exp - xL3xEtOH - (11) nsamples
dt RT K3

[ ( )( )]
5. Discussion
dxEtOH E A1 xL1ExW
- ) wcat k10 exp - xL1xEtOH - + To illustrate the validity of the kinetic model, composition
dt RT K1 profiles were generated from eqs 6-13 for the entire range of

[ (
wcat k20 exp -
E A2
RT )(
xL2xEtOH -
xL2ExW
K2
+ ) 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
E A3 xL3ExW wt % lactic acid, predictions for component profiles were in
k30 exp - xL3xEtOH - (12) reasonably good agreement with the experimental values; for
RT K3
20 wt % lactic acid, predictions were in excellent agreement

[ ( )( )]
with experiments. Error values or deviations (expressed as a
dxW EA1 xL1ExW percentage) between experimental and model predictions for
- ) wcat k10 exp - - xL1xEtOH + each component are provided in Table 3. For experiments with
dt RT K1

[ ( )(
88 wt % feed solution, errors were small during the initial 180

wcat k20 exp -


EA2 xL2ExW
RT K2
- xL2xEtOH + ) 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,

k30 exp - ( )(
EA3 xL3ExW
RT K3
- xL3xEtOH + wcat k40 exp -
E A4
RT
×
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

xL2xW -
xL12
K4
+ k50 exp - ( )(
E A5
RT
xL3xW +
xL1xL2
K5
(13)
to the experiment (Figure 3).
The kinetic model described here is based on the concentra-
tion 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

EtOH:LA catalyst reaction Percent Error


molar ratioa loading (wt %) temperature (°C) L1 L2 L3 L1E L2E L3E EtOH W
88 wt % Lactic Acid Solution
1 3 80 8.7 19.9 24 20.5 37.1 25.1 10.9 5.2
2 3 80 8.9 17.2 15.5 10.1 15.9 28 4.4 2.3
3 3 80 13.9 15.8 11.2 7.7 12.2 20.5 2 2.1
4 3 80 22.2 21 12.8 2.7 16.9 25 1.1 0.9
3 1 80 11.3 9.8 9.3 11.2 16 31.2 1.6 2.1
3 2 80 7.4 10.1 7.6 7 9 29.8 1.4 1.6
3 4 80 15.7 7 4.8 3.6 5.7 29.1 1.4 1
3 5 80 11.5 33.4 27.9 2.6 6.6 16 1.2 0.6
3 3 62 8.3 1.2 1.8 2.5 9.6 31.5 0.8 0.8
3 3 72 5.5 3.5 3.2 0.6 3.6 13.1 0.6 0.2
3 3 90 9.5 41.8 25.9 6.6 14.5 29.6 3 4.2
50 wt % Lactic Acid Solution
1 3 80 2.6 6.9 20.2 49 6.2 0.4
2 3 80 2.2 5.1 11.2 29.2 2.9 0.04
3 3 80 5.1 10.3 4.8 37.1 1.1 0.4
4 3 80 8.4 20.4 5 25.9 0.5 0.6
3 1 80 4.2 1.8 13 14.7 2.0 0.2
3 2 80 5.6 6 4 24.4 1 0.4
3 4 80 6.3 14 3.5 24.7 1.4 0.4
3 3 61 6.6 3.2 14.2 39.1 0.4 0.8
3 3 71 7.1 8.4 5.2 14.8 1.2 0.6
3 3 90 1 20.8 4.1 34.8 1.2 0.5
20 wt % Lactic Acid Solution
1 3 80 1.5 9.2 0.8 0.06
2 3 80 1.4 3.5 0.3 0.05
3 3 80 1.8 4.2 0.2 0.07
4 3 80 4.4 17 0.1 0.2
3 1 80 5.8 9.7 1.3 0.4
3 2 80 1 5.8 0.2 0.03
3 4 80 2 14.1 0.7 0.2
3 3 60 1.2 2.8 0.2 0.08
3 3 70 2.5 5.6 0.5 0.2
3 3 90 1.2 4.6 0.2 0.05
a Molar ratio of ethanol to lactic acid.

that species activity is a fundamentally more appropriate variable water is available for absorption onto resin particles at high lactic
for kinetic modeling; however, for this system, there is little or acid concentration. This hypothesis was further verified by an
no experimental data available on liquid-phase activity coef- experiment in which a known composition of 88 wt % lactic
ficients of lactic acid, its oligomers, and their esters in ethanol- acid and ethanol was brought in contact with the resin at room
water solutions. Prediction of such activity coefficients, which temperature, equilibrated for 30 min, and analyzed. At this
is possible using UNIFAC or other group contribution methods, temperature, no reaction was observed and ethanol was observed
is subject to substantial uncertainties, particularly for the high to be preferentially adsorbed.
concentrations that have been examined here. In fact, we did
prepare an activity-based kinetic model for the lactate esterifi- 6. Conclusions
cation system that included the application of UNIFAC for The kinetics of esterification of lactic acid and its linear
activity coefficient prediction. The optimized activity-based oligomers with ethanol have been described using simple nth-
model gave errors larger than those obtained with the mole order reversible rate expressions. Pre-exponential factors, activa-
fraction-based model; thus, we do not believe there is any tion energies, and equilibrium constants were either fit to the
advantage to application of such a model at this time. experimental data using nonlinear regression analysis or, in some
The choice of simple nth-order kinetic expressions for cases, were calculated directly or taken from other work.
esterification, as opposed to a more mechanistically based Activation energies in the range of 48 000-74 100 kJ/kmol and
approach (such as a Langmuir-Hinshelwood model), was made analysis of mass-transport resistances indicate that the reaction
partially because of uncertainty in the liquid-phase environment is kinetically controlled. The predicted time-dependent species
within the Amberlyst 15 cation-exchange resin. When placed profiles in esterification were in good agreement with experi-
into aqueous solutions that contain ethanol and organic acid, mental results over a wide range of reaction conditions.
Amberlyst 15 swells, because of the selective absorption of
water.22 We observe the same behavior with dilute solutions of Acknowledgment
lactic acid (20 wt % and 50 wt %) in esterification, where water
Financial support from National Corn Growers Association
is present in large excess, relative to lactic acid (20-fold excess and the U.S. Department of Energy are greatly appreciated. The
for 20 wt % and 6-fold excess for 50 wt %), and, thus, free authors also appreciate assistance from MSCRTSF at Michigan
water can absorb. When concentrated (88 wt %) lactic acid, State University for mass spectroscopic analysis and identifica-
where the molar ratio of water to lactic acid is 0.7, is subject to tion of various components.
esterification, ethanol rather than water is preferentially absorbed
into Amberlyst 15. We attribute this behavior to the association Supporting Information Available: Figures containing
of free water with lactic acid, such that very little or no free 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 (8) Beckmann, A.; Nierlich, F.; Popeken, T.; Reusch, D.; von Scala,
free of charge via the Internet at http://pubs.acs.org. 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.
Nomenclature and Units Autocatalyzed and Ion-Exchanged-Resin-Catalyzed Esterification Kinetics
of Lactic acid with Methanol. Ind. Eng. Chem. Res. 2002, 41, 512.
Cj ) concentration of species j in liquid-phase solution (g- (10) Sanz, M. T.; Murga, R.; Beltran, S.; Cabezas, J. L.; Coca, J. Kinetic
mol/kgsoln) Study for Reactive System of Lactic Acid Esterification with Methanol:
De,j ) effective diffusivity of species j in resin (m2/s) Methyl Lactate Hydrolysis Reaction. Ind. Eng. Chem. Res. 2004, 43, 2049.
EAi ) energy of activation for the ith reaction (kJ/kmol) (11) Choi, J. I.; Hong, W. H.; Chang, H. N. Reaction Kinetics of Lactic
ki0 ) pre-exponential factor for the ith reaction (kgsol/kgcat/s) Acid with Methanol Catalyzed by Acid Resins. Int. J. Chem. Kinet. 1996,
28, 37.
kSL,j ) mass-transfer coefficient of species j (m/s)
(12) See, Y.; Hong, W. H.; Hong, T. H. Effects of Operation Variables
Ki ) equilibrium constant for the ith reaction on the Recovery of Lactic Acid in a Batch Distillation Process with Chemical
Lt ) resin particle size equivalent; Lt ) D/6 ) 1 × 10-4 m Reactions. Korean J. Chem. Eng. 1999, 16, 556.
R ) gas constant; R ) 8.31 kJ/(kmol K) (13) Yadav, G. D.; Kulkarni, H. B. Ion-Exchange Resin Catalysis in
rj ) initial reaction rate of species j (g-mol/kgsoln/s) Synthesis of Isopropyl Lactate. React. Funct. Polym. 2000, 44, 153.
T ) temperature (K) (14) Tanaka, K.; Yoshikawa, R.; Ying, C.; Kita, H.; Okamoto, K.
Application of Zeolite T Membrane to Vapor-Permeation-Aided Esterifi-
wcat ) catalyst concentration (kgcat/kgsoln) cation of Lactic Acid with Ethanol. Chem. Eng. Sci. 2002, 57, 1577.
xj ) mole fraction of species j in liquid-phase solution (15) Benedict, D. J.; Parulekar, S. J.; Tsai, S. Esterification of Lactic
Acid and Ethanol with/without Pervaporation. Ind. Eng. Chem. Res. 2003,
Greek Letters 42, 2282.
ηφ2 ) observable modulus (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.
Literature Cited (17) Zhang, Y.; Ma, L.; Yang, J. Kinetics of Esterification of Lactic
Acid with Ethanol Catalyzed by Cation-Exchange Resin. React. Funct.
(1) Asthana, N.; Kolah, A.; Vu, D.; Lira, C. T.; Miller, D. J. A
Polym. 2004, 61, 101.
Continuous Reactive Separation Process for Ethyl Lactate Formation. Org.
Proc. Res. DeV. 2005, 9, 599. (18) Rohm & Hass. Product Data Sheet, Amberlyst-15 Dry. Available
(2) Asthana, N.; Kolah, A.; Lira, C. T.; Miller, D. J. Improved Process via the Internet at http://www.rohmhaas.com/ionexchange/IP/sac.htm.
for Production of Organic Acid Esters. U.S. Patent Application No. 2006/ (19) Reid, R. C.; Prausnitz, M. J.; Sherwood, T. K. The Properties of
0014977 A1, 2004. Gases and Liquids; McGraw-Hill: New York, 1987.
(3) Vu, D. T.; Kolah, A. K.; Asthana, N. S.; Peereboom, L.; Lira, C. (20) Chakrabarti, A.; Sharma, M. M. Cationic Ion-Exchange Resin as
T.; Miller, D. J. Oligomer Distribution in Concentrated Lactic acid Solutions. Catalyst. React. Polym. 1993, 20, 1.
Fluid Phase Equilib. 2005, 236, 39. (21) Gangadwala, J.; Mankar, S.; Mahajani, S.; Kienle, A.; Stein, E.
(4) Holten, C. H. Lactic Acid: Properties and Chemistry of Lactic Acid Esterification of Acetic Acid with Butanol in the Presence of Ion-Exchange
and DeriVatiVes; Verlag Chemie: Copenhagen, 1971. Resins as Catalysts. Ind. Eng. Chem. Res. 2003, 42, 2146.
(5) Taylor, R.; Krishna, R. Modelling Reactive Distillation. Chem. Eng. (22) Mazzotti, M.; Neri, B.; Gelosa, D.; Kruglov, A.; Morbidelli, M.
Sci. 2000, 55, 5183. Kinetics of Liquid-Phase Esterification Catalyzed by Acidic Resins. Ind.
(6) Popken, T.; Gotze, L.; Gmehling, J. Reaction Kinetics and Chemical Eng. Chem. Res. 1997, 36, 3.
Equilibrium of Homogeneously and Heterogeneously Catalyzed Acetic Acid
Esterification with Methanol and Methyl Acetate Hydrolysis. Ind. Eng. ReceiVed for reView December 6, 2005
Chem. Res. 2000, 39, 2601. ReVised manuscript receiVed April 18, 2006
(7) Dhale, A. D.; Myrant, L. K.; Chopade, S. P.; Jackson, J. E.; Miller, Accepted May 10, 2006
D. J. Propylene Glycol and Ethylene Glycol Recovery from Aqueous
Solutions via Reactive Distillation. Chem. Eng. Sci. 2004, 59, 2881. IE0513604

Вам также может понравиться