You are on page 1of 9

4250

Ind. Eng. Chem. Res. 1999, 38, 4250-4258

Kinetic Model of Linear Complex Esterification between


2-Butyl-2-ethyl-1,3-propanediol, Adipic Acid, and Octanoic Acid
Matti O. Lehmus, Sami Toppinen, Maaria K. Sela1 ntaus, Nina M. Kopola, and
A. Outi I. Krause*,

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Fortum Oil and Gas Oy, P.O. Box 100, FIN-00048 Fortum, Finland, Technology Center, Neste Chemicals Oy,
P.O. Box 310, FIN-06101 Porvoo, Finland, and Department of Chemical Technology, Helsinki University of
Technology, P.O. Box 6100, FIN-02015 HUT, Finland

In the present study, a kinetic model of the autocatalyzed linear complex esterification of 2-butyl2-ethyl-1,3-propanediol (BEPD), octanoic acid, and adipic acid was developed. The presented
model describes complex esterification kinetics at low to intermediate conversions. For the first
time, the concentrations of all components involved in the complex esterification reaction system
were modeled individually, and the model was able to estimate accurately all 10 analyzed
component concentrations in the entire temperature range studied. The model utilizes a simple
second order rate law derived from the esterification mechanism to describe esterification kinetics.
In order to avoid overparametrization of the model, simplifying assumptions regarding
substitution effects were made which allowed the description of the entire set of 125 plausible
esterification reactions with six kinetic parameters. The model was applied to both the simple
polyol esterification of BEPD and octanoic acid and the complex esterification involving
additionally adipic acid, and all esterifications were carried out in the temperature range of
170-190 C. As kinetic parameters of the simple polyol esterification could also be used in the
modeling of the complex esterification, only four kinetic parameters were needed to be fitted
simultaneously to describe complex esterification kinetics.
Introduction
Complex esters are oligoesters formed in the reaction
of a polyfunctional alcohol, a dicarboxylic acid, and a
monofunctional alcohol or acid. The reaction product of
complex esterification is a mixture of simple mono- and
diesters and oligoesters of higher molecular weight. The
chemical composition of the reaction product depends
on various factors such as reactant stoichiometry and
reaction conditions. In order to tailor complex esters for
specific applications, it is necessary to understand and
control the kinetics of the complex reaction system.
Very few research results on complex esterification
kinetics have been published, whereas the kinetics of
polyesterifications and polyol esterifications has received much attention. As early as 1939, Flory proposed
a third order rate law to describe polyesterification
kinetics.1 However, already Flory noticed that a third
order rate law does not describe the polyesterification
kinetics accurately at low acid conversions. It has been
proposed that an accurate kinetic model should take into
account not only hydrolysis reactions but also changes
of the esterification mechanism with increasing conversion caused e.g. by the decreasing polarity of the
reaction medium. As reviewed by Fradet and Marechal
(1982),2 several attempts to develop a rate expression
valid over the entire conversion range have been made,
e.g. Tang and Yao (1959),3 Fang et al. (1975),4 Lin and
Hsieh (1977),5 and Chen and Hsiao (1981).6 More
* Author to whom correspondence should be addressed.
Telephone: +358-9-4512613. Fax: +358-9-4512622. E-mail:
Krause@polte.hut.fi.
Fortum Oil and Gas Oy.
Neste Chemicals Oy.
Helsinki University of Technology.

recently e.g. Bacaloglu et al. (1988)7 and (1998),8 Paatero


et al. (1994),9 Salmi et al. (1994),10 and Lehtonen et al.
(1996)11 have developed kinetic models of polyesterification. In spite of all the efforts, some controversy still
exists about the exact reaction mechanism and the rate
law of autocatalyzed (poly)esterification. Since the esterification reaction apparently proceeds simultaneously
via several mechanisms, the exact rate law becomes
very complex.
In the kinetic modeling of polyesterifications, substitution effects have usually been neglected and the
principle of equal reactivity of functional groups introduced by Flory1 has been applied. However, especially
for low degrees of polymerization the accuracy of this
approximation is not evident. In kinetic studies of
simple neopentyl polyol esterifications it has been found
necessary to observe substitution effects in the determination of the esterification rates of subsequent hydroxyl groups of the same polyol molecule, since the
reactivity of functional groups is affected by substituents
introduced near to them.12-14 Kinetic models of polyesterification taking into account substitution effects have
been developed e.g. by Gupta et al. (1979),15,16 who
introduced two different rate constants to account for
the different reactivity of monomers and higher oligomers. The value of the rate constant describing the
reaction rate of the esterification between monomer and
oligomer was assumed to be between these two limiting
values.
A reaction system resembling complex esterification
has been studied recently by Lopez-Gonzalez et al.
(1998),17 who developed a kinetic model for the copolyesterification between o-phthalic anhydride, oleic acid,
and neopentyl glycol. The concentrations of eight glycolic
structural units could be determined by NMR tech-

10.1021/ie980633a CCC: $18.00 1999 American Chemical Society


Published on Web 10/09/1999

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999 4251

niques, and their time dependence could be modeled


with 12 kinetic parameters, four of which had to be fixed
during the parameter estimation procedure.
In the present work, a kinetic model describing the
concentrations of individual components in a complex
esterification reaction system was developed for the first
time. The reaction product of complex esterification
consists of a mixture of simple polyol esters and complex
esters which differ in their chemical and physical
characteristics. Consequently, the modeling of carboxylic
acid or hydroxyl group concentrations as usually carried
out in kinetic studies of polyesterifications is not sufficient to characterize complex esterification reaction
products.
However, componentwise modeling leads to an increased number of parameters in the kinetic model.
Depending on the number of analyzed component
concentrations, only a limited number of parameters can
be determined unambiguously and reliably. Since not
all individual components could be analyzed quantitatively although sophisticated analysis methods were
applied, simplifying assumptions had to be applied in
order to reduce the number of kinetic parameters in the
model. By limiting the conversion range to low and
intermediate conversions, some important approximations regarding the rate law could be utilized. A more
sophisticated rate law with more kinetic parameters can
be incorporated, when progress in the development of
analysis methods allows the quantitative determination
of more component concentrations. Consequently, this
study represents a first step toward the target of
developing a kinetic model which is able to describe
complex esterification kinetics accurately in the entire
conversion range.
Experimental Section
Two series of kinetic experiments were conducted at
three temperatures. In the first series, the simple polyol
esterification between 2-butyl-2-ethyl-1,3-propanediol
(BEPD) and octanoic acid (OA) was studied, whereas
in the second series adipic acid (AA) was also added to
the system.
Apparatus. The experiments were carried out in a
1000 cm3 tank reactor equipped with a two-blade stirrer,
three baffle plates, and a cooling condenser allowing the
withdrawal of water during the esterification. In order
to facilitate water withdrawal, nitrogen was continuously fed through the reactor in addition to which the
esterifications were carried out at a reduced reaction
pressure of 780 mbar. The reactor was heated with an
oil-filled jacket, and the temperature in the reactor was
controlled automatically by adjustment of the heating
oil temperature according to reactor temperature measurement.
BEPD was placed in the reactor vessel prior to the
experiment and heated to the desired temperature. The
carboxylic acid reactants were heated in a separate
vessel to the same temperature, and esterification
reactions were started by feeding the acid into the
reactor. The samples withdrawn from the reaction
mixture were immediately cooled with solid carbon
dioxide to avoid consecutive reactions.
Analysis. The samples withdrawn during simple
esterifications of BEPD and octanoic acid were analyzed
with a Hewlett-Packard 6890 gas chromatograph (GC)
equipped with a capillary column (J&W DB-1, 60 m)
and a flame ionization detector (FID). GC analysis was

also used to determine the BEPD and octanoic acid


contents of the complex esterification samples, and the
analyses were performed with a Hewlett-Packard 5890
gas chromatograph equipped with a J&W DB-WAX
column (30 m) and a FID. Response factors were
determined with calibration solutions.
The adipic acid content was determined with a
gradient elution high-performance liquid chromatography (HPLC) technique, using a mixture of water and
acetonitrile as solvent. The apparatus used was a HP
1090 liquid chromatograph equipped with a Merck
Lichrosorb RP-18 column. Elution of adipic acid was
detected by UV absorption at 210 nm with a diode array
detector. The reaction products of complex esterification
were analyzed by isocratic HPLC analysis with a
modular Waters chromatograph equipped with a Merck
Lichrosorb RP-18 column and a refractive index detector. A 98:2 vol % mixture of methanol and water was
utilized as solvent, and a column temperature of 35 C
was applied. The chromatogram displayed two sets of
peaks which were assigned to two different homologous
series of octoate single and double end-capped complex
esters identified by field desorption mass spectrometric
(FD-MS) analysis. The homologous series can be described by formulas OA-(BEPD-AA)n-BEPD-OA (n )
0...4) and OA-(BEPD-AA)m-BEPD (m ) 0...1). As proposed by Robertson et al. (1993),18 complex esters of the
same homologous series were assumed to elute in the
order of increasing degree of polymerization. Calibration
was carried out by using the response factors of BEPD(mono)octoate and BEPD-dioctoate for the single and
double end-capped species, respectively.
Chemicals. The following chemicals were used in the
experiments: octanoic acid (>98 wt %, Fluka Chemika
AG), adipic acid (>99 wt %, Fluka Chemika AG), and
BEPD (99.5 wt %, Neste Oxo AB).
Simple and Complex Esterification Experiments. Both sets of experiments were carried out at
reaction temperatures of 170, 180, and 190 C and with
a stirrer speed of 500 rpm. Experiments conducted at
higher stirrer speeds implied that mass transfer resistance did not affect reaction rates. In both simple and
complex esterifications, hydroxyl and carboxylic acid
group amounts were equimolar. In complex esterification experiments, the molar ratio of octanoic acid and
adipic acid was 89:11.
Results and Discussion
Kinetic Model of Complex Esterification. (a)
Reactions. In order to systemize the treatment of
different esterification reactions, they were grouped into
four categories, as shown in Table 1. The group of simple
esterifications encompasses the mono- and diesterification reactions of BEPD and octanoic acid or adipic acid,
whereas the second group titled chain growth reactions
describes all reactions in which BEPD or adipic acid
monomers add to a BEPD-adipate chain. The third
group of complex esterifications includes all reactions
in which esters containing both acids are formed. In
addition to the stepwise addition of monomers described
in the first three groups, the fourth group of oligomer
esterifications was incorporated to account for esterifications in which neither reactant is a monomer.
Since it was necessary to reduce the number of kinetic
parameters, two approximations regarding substitution
effects were made. First of all, it was assumed that the
first monomer of the substituent determines the sub-

4252

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999

Table 1. Esterification Reactions Incorporated into the


Kinetic Model
Simple Esterifications
ke1

BEPD + OA 98 BEPD-OA + H2O


ke2

BEPD-OA + OA 98 OA-BEPD-OA + H2O


ke3

BEPD + AA 98 BEPD-AA + H2O


ke4

BEPD-AA + BEPD 98 BEPD-AA-BEPD + H2O


ke5

BEPD-AA + AA 98 AA-BEPD-AA + H2O

(1)
(2)
(3)
(4)
(5)

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Chain Growth Reactions


kchain1,n

(AA-BEPD)n + AA 98 (AA-BEPD)n-AA + H2O


kchain2,n

BEPD-(AA-BEPD)n + AA 98 (AA-BEPD)n+1 + H2O


kchain3,n

(BEPD-AA)n + BEPD 98 BEPD-(AA-BEPD)n + H2O


kchain4,n

AA-(BEPD-AA)n + BEPD 98 (AA-BEPD)n+1 + H2O

(6)
(7)
(8)
(9)

Complex Esterifications
kcomplex1,i

Ri-AA-BEPD + OA 98 Ri-AA-BEPD-OA + H2O


kcomplex2

OA-BEPD + AA 98 AA-BEPD-OA + H2O

(10)
(11)

kcomplex3,n

OA-BEPD-(AA-BEPD)n + AA 98
OA-(BEPD-AA)n+1 + H2O (12)

intermediate conversion kinetics can be modeled reasonably accurately in spite of discarding hydrolysis
reactions if reaction conditions and equipment are
chosen to optimize water withdrawal from the reaction
mixture. In the developed kinetic model, esterification
reactions were consequently treated as irreversible,
which considerably reduces the number of kinetic
parameters to be determined. Since exchange reactions
require the addition of a catalyst or elevated temperatures (>250 C) to proceed, they were omitted from the
model as well. Side reactions could also be discarded as
GC analysis of the distillate did not indicate the
formation of any side reaction products.
The kinetics of the complex reaction system could
consequently be described with six reaction rate constants ke1-ke6. The number of components and reactions
which have to be observed in the kinetic model depends
on the maximal degree of polymerization DPmax. HPLC
analysis indicated that the reaction mixture did not
contain important amounts of complex esters with a DP
exceeding 11. All possible components with a DP not
exceeding 11 and all esterification reactions in which
these components are formed were thus incorporated
into the model, making up a total of 33 components and
125 esterification reactions.
(b) Rate Equation. Esterifications carried out in the
absence of an added catalyst most commonly take place
by two mechanisms, a noncatalyzed and an autocatalyzed mechanism.2 The noncatalyzed bimolecular mechanism (BAc-2) yields an esterification rate law which
is first order in both acid and alcohol:

re,noncatalyzed ) ke,noncatalyzedcRCOOHcROH

(20)

kcomplex4,n

OA-(BEPD-AA)n + BEPD 98
BEPD-(AA-BEPD)n-OA + H2O (13)

stitution effect. This approximation yields the following


simplifications

The autocatalytic mechanism (AAc-2), on the other


hand, proceeds via protonation of the carboxylic acid by
another carboxylic acid and a subsequent nucleophilic
addition of an alcohol to the carbenium ion formed in
the protonation reaction.2 Since the nucleophilic addition of an alcohol ROH to the protonated acid RiC+(OH)2 is considered the rate determining step of the
esterification mechanism, the general rate law can be
expressed as

kchain1,n ) kchain2,n ) kcomplex3,n ) ke5

(15)

re,autocatalyzed ) kRDcRiC+(OH)2cROH

kchain3,n ) kchain4,n ) kcomplex4,n) ke4

(16)

Oligomer Esterifications
koligomer,ij

Ri-BEPD + Rj-AA 98 Rj-BEPD-AA-Rj + H2O

(14)

Second, it was approximated that the substitution


effects of adipate and octoate groups are identical since
the effect of the second carboxylic acid group of the
adipate group on the the unreacted hydroxyl group is
weakened by the long molecular chain situated between
them. On the basis of this additional approximation, the
following relationships are obtained

kcomplex1,i ) ke2

(17)

kcomplex2 ) ke5

(18)

koligomer,ij ) constant ) ke6

(19)

The significance of hydrolysis reactions becomes dominant at high conversions when the concentration of
esters in the reaction mixture is high and the reaction
is thermodynamically controlled. Consequently, low and

(21)

Paatero et al.9 have proposed that protonation of the


reacting acid RiCOOH may follow two paths depending
on the acid conversion. At low or intermediate conversions separate ions are formed according to
K1,ij

RiCOOH + RjCOOH 798 RiC+(OH)2 + RjCOO- (22)


whereas ion pairs are formed at high conversions as
shown in
K2,ij

RiCOOH + RjCOOH 798 (RiC+(OH)2RjCOO-)

(23)

Since complex esterification involves numerous carboxylic acids which may function as proton donators in
reactions 22 and 23, RjCOOH may be any carboxylic
acid in the reaction mixture (including also RiCOOH).

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999 4253

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Application of the quasi-equilibrium approximation to


reactions 22 and 23 yields
K1,ij ) c+
i cj /cicj

(24)

K2,ij ) cij/cicj

(25)

where ci and cj are the concentrations of acids RiCOOH


and RjCOOH, c+
i and cj are the concentrations of the
+
protonated acid RiC (OH)2 and the carboxylate anion
RjCOO-, respectively, and cij is the concentration of the
ion pair formed in reaction 23. Since the studied
temperature range was narrow, protonation reaction
equilibrium constants were approximated to be temperature-independent. The rigorous thermodynamic
interpretation of the reaction mechanism implies that
equilibrium constants are functions of the reaction
milieu, as discussed e.g. by Fang et al.4 However, in
order to avoid overparametrization of the model, the
conversion dependence of the equilibrium constants was
omitted as well and the following approximation was
applied:

K1,ij ) constant ) K1

(26)

K2,ij ) constant ) K2

(27)

Equations 26 and 27 also imply that the values of the


protonation equilibrium constants do not depend on the
carboxylic acids involved in the protonation reaction.
The accuracy of this approximation depends primarily
on the structural and chemical similarity of the reacting
acids. Inserting eqs 26 and 27 in eqs 24 and 25 and
taking into account all possible protonation reactions
of RiCOOH yields

c+
i
K1 )

j c-j

ci

K2 )

j cj

j cij
ci

j cj

(28)

(29)

(30)

Combining eqs 28 and 30 and summing over all different


acids RiCOOH allows the deduction of equation

j c+j ) K11/2j cj

(31)

Inserting (30) and (31) in (28) yields


1/2
c+
i ) K1 c i

(32)

The concentration of all ion pairs of RiCOOH (i.e. jcij)


can be obtained from (29):

j cij ) K2cij cj

re-autocatalyzed,i ) ke1,autoK11/2cRiCOOHcROH +
ke2,autoK2cRiCOOH

j cR COOHcROH
j

(34)

Combining with the rate equation (20) for the noncatalyzed esterification yields the complete rate equation

re,i ) (ke1,autoK11/2 + ke,noncatalyzed)cRiCOOHcROH +


ke2,autoK2cRiCOOH

j cR COOHcROH
j

(35)

Since the effect of ion pairs on the reaction mechanism


becomes important only at high conversions, an approximate rate law applicable for low and intermediate
conversions can be deduced by discarding ion pair
formation as demonstrated by Lehtonen et al.,11 i.e. by
assuming K2 0 in the conversion range studied. Thus
following the second order rate law

re,i ) kecRiCOOHcROH

(36)

is finally obtained, with

ke ) ke1,autoK11/2 + ke,noncatalyzed

(37)

Reaction rate constants ke were defined relative to


functional group concentrations, which implies doubling
of the reactant concentration in case it contains two
equivalent functional groups.
The total rate of formation ri of component i is
obtained by observing its rate of formation in all
individual esterification reactions n
125

ri )

(i,nre,n)

(38)

n)1

Application of the principle of electric neutrality implies

j c+j ) j c-j

Inserting eqs 32 and 33 into eq 21 finally yields the rate


law of the autocatalytic esterification

(33)

where i,n represents the stoichiometric coefficient of


component i in reaction n. Since the number of possible
esterification reactions is large and depends on the
reaction conditions, eq 38 for all components was formed
with a computerized algorithm which takes into account
the maximal degree of polymerization observed.
In the computerized algorithm each component is
unambiguously defined by three numbers {X Y Z}. X
and Y denote the number of unreacted hydroxyl and
carboxylic acid end groups, respectively, and Z is the
chain length of the molecule (e.g. octanoic acid is
consequently defined as {0 1 1} and OA-(BEPD-AA)4
as {0 1 9}). The algorithm is visualized in the flow chart
presented in Figure 1. After the definition of the
maximal degree of polymerization in step 1, all plausible
esterification reactions are formed in steps 2 and 3. The
reactions are divided into six groups according to the
reaction rate constant involved, and the plausible
reactants {X Y Z} belonging to each group are generated
in a loop algorithm (step 2). The products of each
reaction are calculated by simple additive and subtractive operations (step 3), after which the rate of each
individual reaction is obtained by applying the general
rate law (36) to the reaction under consideration (step
4). For each reaction, the reaction rate rreaction is added
to the rate of formation of the product i.e. rproduct.
Accordingly, -rreaction is added to the rates of formation

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

4254

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999

Figure 1. Flow chart of the algorithm generating the rates of formation ri for each component.

of the reactants (step 5). Subprograms which calculate


molecular masses and double the concentrations of
components with two equivalent functionals have been
omitted from the flow chart in Figure 1.
(c) Reactor Model. The differential mole balances
for each component i served as a starting point for the
modeling of the reactor:

dni/dt ) -n i,distillate + mri

(39)

where ni denotes the molar amount of component i in


the system and n i,distillate represents the molar flow of
component i leaving the system in the distillate. The
rate of formation of component i was defined with
respect to the mass m of the reaction mixture. Due to
the discrete nature of sampling, sample removal is not
observed in the differential mole balances (39) and was
consequently taken into account by direct subtraction
from differential mole balance solutions. The mass of
the reaction mixture m(t) was calculated from the mole
balances as follows:

m(t) )

i ni(t)Mi

(40)

where Mi denotes the molar mass of component i. The


molar flow of component i in the distillate can be
expressed as

distillateci,distillate
n i,distillate ) m

(41)

with m
distillate denoting the mass flow rate of the distillate. Since the distilled organic compounds were scarcely
soluble in the aqueous phase, the organic and aqueous
phases of the distillate could be modeled separately. The
approximate concentration of compound i in the organic
phase was obtained by assuming that the composition
of the organic distillate remained constant during
esterification and could consequently be determined by
GC analysis at the end of the experiment. Since the
organic distillate contained more than 80 wt % octanoic
acid, this approximation does not cause any significant
errors in the mole balances. Distillate formation was
monitored individually for the aqueous and organic
phases by measuring the volume of both distilled phases
when withdrawing the samples. Resulting from the
approximation introduced above, the average densities

Table 2. Weighting Factors and Average Concentrations


of Analyzed Components
components

pi

cav/(wt %)

OA, BEPD, BEPD-OA, OA-BEPD-OA


1 10...50
AA, BEPD-AA-BEPD-OA, OA-BEPD-AA-BEPD-OA
10 1...7
OA-BEPD-(AA-BEPD)n-OA (n ) 2...4)
100 0.1...1.0

of the phases were assumed to remain constant during


esterification. Determination of the average density of
both phases at the end of the experiments thus allowed
the calculation of data points describing distillate mass
formation. Both organic and aqueous distillate masses
were modeled with a cubic spline function, i.e. by fitting
a third order polynomial function between each pair of
successive data points. The mass flow rates of both
distillate phases were calculated by differentiating the
obtained spline functions.
(d) Parameter Estimation Procedure. The kinetic
model of complex esterification consisted of 33 ordinary
differential equations (ODEs) obtained from the differential mole balances (39) for each component. During
parameter estimations, the ODEs were solved with the
ODE-solution subroutine LSODE (Hindmarsh),19 which
uses the backward difference method (Gear).20 The
kinetic parameter values were fitted by non-linear
regression analysis. The objective function Q for the
parameter estimation was defined as follows:

Q)

i t pi(wi,exptl(t) - wi,estd(t))2

(42)

where i refers to analyzed component i and t to sampling


time, pi is a weighting factor and wexptl and westd denote
the experimentally measured and the estimated weight
fractions, respectively. Estimated weight fractions were
calculated from the numerical solutions of the kinetic
model according to eq 43.

wi,estd(t) ) ni(t)Mi/m(t)

(43)

The application of weighted concentrations in the objective function improves the accuracy of the fit for lowconcentration components. The approximate equality of
the weighted concentration values was used as the
criterion for the determination of suitable weighting
factors. Weighting factor values utilized in the parameter estimation procedure are summarized in Table 2.

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999 4255
Table 3. Kinetic Parameters for the Esterification of BEPD and Octanoic Acid
data sets

T/C

ke1 105/(kg mol-1 s-1)

ke2 105/(kg mol-1 s-1)

separate

170
180
190
170a
180
190a

7.2 ( 0.4
10.2 ( 0.7
13.8 ( 0.9
7.2
10.1 ( 0.4
13.9

5.1 ( 0.2
7.3 ( 0.3
10.1 ( 0.3
5.1
7.2 ( 0.1
10.1

31.8
39.4
26.8

55.8

58.7

98.4

56.0 ( 7.4

58.6 ( 4.0

merged

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

aValues

Ea1/(kJ mol-1)

Ea2/(kJ mol-1)

calculated with Arrhenius equation (44).

Figure 2. Arrhenius plots of mono- and diesterification rate


constants for esterification of BEPD and octanoic acid.

The objective function was minimized with the Levenberg-Marquardt method (Marquardt)21 using the
subroutine NL2SOL (Dennis et al.)22 coupled with the
ODE-solver LSODE.
Parameter Estimation Results. The simple esterification between BEPD and octanoic acid was modeled
by fitting the kinetic model with two parameters ke1ke2 to the experimental data. The modelling was initially
carried out by fitting the kinetic parameters separately
at each temperature. The Arrhenius plots shown in
Figure 2 indicate that the temperature dependence of
the rate constants is accurately described by the Arrhenius equation

k(T) ) kTref exp

((

Ea 1
1
R Tref T

))

(44)

This result supports the validity of the approximation


according to which the protonation equilibrium constants incorporated into the rate constants are temperature-independent. In a second step, parameter estimation was thus carried out by merging the data sets
obtained at different temperatures and incorporating
the Arrhenius equation with Tref ) 453.15 K into the
model. The model was able to describe accurately
esterification kinetics over the entire temperature range
studied independent of the parameter estimation strategy. An example of the obtained fit for the esterification
carried out at 170 C is presented in Figure 3. Results
for the reaction rate constants and the activation
energies including 95% confidence intervals are summarized in Table 3. It can be seen that merging of the
data sets improves estimation statistics slightly but does
not affect the estimated values significantly. The estimation statistics is very satisfactory, and the determined activation energies correspond to typical values
reported for esterification reactions. A distinctive negative substitution effect of the octoate group with ke2/ke1
) 0.7 is observed.

Figure 3. Experimental and estimated concentrations of octanoic


acid (OA), BEPD, BEPD-octoate (BO), and BEPD-dioctoate (OBO)
for the esterification of BEPD and octanoic acid carried out at 170
C (experimental, dots; estimated, continuous lines).

Complex esterification modeling was initiated by


fitting all six reaction rate constants ke1-ke6 simultaneously to the experimental data obtained at 180 C.
Since the values obtained for ke1 and ke2 did not differ
significantly from the values determined previously for
the simple esterification system, it was concluded that
the addition of adipic acid does not affect significantly
the rate constants of octanoic acid esterification. This
result supports the validity of approximation 26, since
mixture effects would be expected if the presence of
adipic acid affected the protonation equilibrium of
octanoic acid. In the second modeling stage, parameter
values determined in simple esterification experiments
were thus used for ke1 and ke2 and only four kinetic
parameters ke3-ke6 were fitted.
The kinetic parameters were first determined separately at each reaction temperature. The kinetic model
was able to describe well the time dependence of all
measured concentrations in the entire conversion and
temperature range studied. Measured and estimated
concentration curves for the complex esterification
conducted at 170 C are presented in Figures 4-6. The
concentration curves for components that were not
analyzed could be simulated as shown in Figure 7 for
some representative components.
The results for the parameter values are presented
in Table 4. The Arrhenius plots presented in Figure 8
indicate that the temperature dependence of the rate
constants can be described quite accurately by the
Arrhenius equation (44). The simplified rate law which
discards hydrolysis reactions and the temperature and
conversion dependence of the equilibrium constants
represesents the most probable cause for the Arrhenius
plots slight deviation from linearity. However, model
accuracy was not impaired by merging of the data sets
obtained at three reaction temperatures and incorporation of eq 44 into the kinetic model, which indicates that
the accuracy of the Arrhenius equation in the descrip-

4256

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999

Table 4. Kinetic Parameters of Complex Esterification between BEPD, Adipic Acid, and Octanoic Acid
ke3 105/(kg mol-1 s-1)

ke4 105/(kg mol-1 s-1)

170
180
190

6.1 ( 2.2
6.2 ( 4.4
7.7 ( 5.7

17.7 ( 3.2
40.3 ( 16.6
72.6 ( 46.2

Ea/(kJ mol-1)

20.1

120.4

T/C

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Separate Data Sets


8.6 ( 4.1
15.5 ( 8.6
19.1 ( 13.4

170a
180
190a

5.7
6.7 ( 2.4
7.8

18.8
37.5 ( 8.5
72.9

Ea/(kJ mol-1)

26.6 ( 67.1

115.9 ( 43.8

ke5 105/(kg mol-1 s-1)

68.2
Merged Data Sets
9.7
14.0 ( 2.9
19.9
61.2 ( 58.8

ke6 105/(kg mol-1 s-1)


6.2 ( 2.0
8.3 ( 4.2
19.6 ( 9.8

Q
68.8
214.0
264.7

97.3
5.4
10.0 ( 4.5
18.2

564.3

103.8 ( 52.9

Values calculated with Arrhenius equation (44).

Figure 4. Experimental and estimated concentrations of octanoic


acid (OA), BEPD, BEPD-octoate (BO), and BEPD-dioctoate (OBO)
for complex esterification carried out at 170 C (experimental, dots;
estimated, continuous lines).

Figure 5. Experimental and estimated concentrations of adipic


acid (AA), BEPD-AA-BEPD-OA (BABO), and OA-BEPD-AA-BEPDOA (OBABO) for complex esterification carried out at 170 C
(experimental, dots; estimated, continuous lines).

tion of the temperature dependence of the rate constants


is satisfactory. The results for the parameter values
obtained for the merged data set are summarized in
Table 4. The obtained values do not differ significantly
from the values determined separately at each reaction
temperature, but estimation statistics for the rate
constants improves slightly. Although the error margins
for the activation energy values are large, the two
different parameter estimation strategies yield similar
values for the activation energies.
As expected, minor discrepancies between estimates
and experimental results can be observed at high
conversion values, which can be explained with the
inability of the simplified rate law to account for
hydrolysis reactions and a change of the reaction
mechanism at high acid conversion.

Figure 6. Experimental and estimated concentrations of OABEPD-(AA-BEPD)2-OA (OB(AB)2O), OA-BEPD-(AA-BEPD)3-OA


(OB(AB)3O), and OA-BEPD-(AA-BEPD)4-OA (OB(AB)4O) for the
complex esterification carried out at 170 C (experimental, dots;
estimated, continuous lines).

Figure 7. Estimated concentrations of AA-BEPD-OA (ABO),


BEPD-(AA-BEPD)2-OA (OB(AB)2), AA-BEPD (AB), and BEPDAA-BEPD (BAB) for complex esterification carried out at 170 C
(experimental, dashes; estimated, continuous lines).

Conclusions
A kinetic model was developed for the complex reaction system of BEPD, adipic acid, and octanoic acid. In
spite of the large number of components involved, the
attempt of modeling the concentrations of all individual
components was made. The following conclusions can
be drawn from the results:
(1) The experimentally determined concentrations of
10 components could be modeled accurately at low and
intermediate conversions in the entire temperature
range studied. Consequently, the approach of modeling
individual component concentrations even for a large
number of components proved fruitful, which indicates
that all important reactions are observed.

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999 4257
R ) general gas constant (J K-1 mol-1)
r ) rate of reaction (mol kg-1 s-1)
t ) time (s)
w ) weight fraction
Subscripts and Superscripts
+, - ) indices referring to RC+(OH)2 and RCOO-, respectively
av ) average
e ) esterification
i, j ) component indices
n ) reaction index
n, m ) indices for repeating units in molecular structure
ref ) reference

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

Figure 8. Arrhenius plots of complex esterification rate constants


ke3-ke6.

(2) The kinetics of the complex reaction system


studied was succesfully described by fitting merely four
parameter values, which indicates that the simplifications incorporated in the model do apply with sufficient
accuracy. The results of this study thus prove that
substitution effects can be modeled by observing only
the first monomer of the substituent and that complex
esterification kinetics at low and intermediate conversions can be described with a simple second order rate
law.
(3) Since the accuracy of the six parameter model is
excellent in spite of the large number of simplifications
incorporated, the introduction of additional kinetic
parameters does not improve model reliability because
determination of unambiguous parameter values becomes problematic. If the quantitative analysis of an
even larger number of complex esters becomes feasible
in the future, a larger number of kinetic parameters can
be determined, which allows the incorporation of a more
detailed rate law into the kinetic model presented in
this study. The change in the esterification mechanism
and the effect of reverse reactions at high conversion
values can then be taken into account, allowing accurate
componentwise modeling even at high conversions.
(4) The approach developed in this work is generally
applicable to the modeling of all complex esterifications,
for which the number of kinetic parameters can be
reduced by grouping similar reactions into entities
described by a small number of parameters. Since monoand dicarboxylic acids were assumed to have similar
substitution effects, the kinetic model is optimally suited
to describe complex esterifications involving structurally
similar mono- and dicarboxylic acids. Additionally, the
model can be applied in the kinetic modeling of any
other linear condensation reaction system with reactants fulfilling the same criterion.
Notation
Ea ) activation energy (kJ mol-1)
K ) equilibrium constant
k ) rate constant (kg mol-1 s-1)
m ) mass (of liquid phase) (kg)
M ) molecular weight (kg mol-1)
n ) amount of substance (mol)
n ) molar flow of substance (mol s-1)
) stoichiometric coefficient
p ) weight factor in objective function
Q ) objective function (residual sum of squares)

Abbreviations
AA, A ) adipic acid
BEPD, B ) 2-butyl-2-ethyl-1,3-propanediol
DP ) degree of polymerization
OA, O ) octanoic acid
ODE ) ordinary differential equation
ROH ) alcohol
RC+(OH)2 ) carbenium ion formed by protonation of
RCOOH
RCOO- ) carboxylate anion
RCOOH ) carboxylic acid

Literature Cited
(1) Flory, P. J. Kinetics of Polyesterification: A Study of the
Effects of Molecular Weight and Viscosity on Reaction Rate. J.
Am. Chem. Soc. 1939, 61, 3334-3340.
(2) Fradet, A.; Marechal, E. Kinetics and Mechanisms of
Polyesterifications I. Reactions of Diols with Diacids. Adv. Polym.
Sci. 1982, 43, 51-142.
(3) Tang, A. C.; Yao, K. S. Mechanism of Hydrogen Ion
Catalysis in Esterification. II. Studies on Kinetics of Polyesterification Reactions between Dibasic Acids and Glycols. J. Polym.
Sci., Polym. Chem. Ed. 1959, 35, 219-233.
(4) Fang, Y.-R.; Lai, C.-G.; Lu, J.-L.; Chen, M.-K. Kinetics and
Mechanism of Polyesterification of Binary Acids and Binary
Alcohols. Sci. Sin. 1975, 18, 72-87.
(5) Lin, C. C.; Hsieh, K. H. The Kinetics of Polyesterification.
I. Adipic Acid and Ethylene Glycol. J. Appl. Polym. Sci. 1977, 21,
2711-2719.
(6) Chen, S.; Hsiao, J. Kinetics of Polyesterification. I. Dibasic
Acid and Glycol Systems. J. Polym. Sci., Polym. Chem. Ed. 1981,
19, 3123-3136.
(7) Bacaloglu, R.; Maties, M.; Csunderlik, C.; Cotarca, L.;
Moraru, A.; Gros, J.; Marcu, N. Kinetics of Polyesterification and
Models, 2. Angew. Makromol. Chem. 1988, 164, 1-20.
(8) Bacaloglu, R.; Fisch, M.; Biesiada, K. Kinetics of Polyesterification of Adipic Acid With 1,3-Butanediol. Polym. Eng. Sci.
1998, 38, 1014-1022.
(9) Paatero, E.; Narhi, K.; Salmi, T.; Still, M.; Nyholm, P.;
Immonen, K. Kinetic Model for Main and Side Reactions in the
Polyesterification of Dicarboxylic Acids with Diols. Chem. Eng. Sci.
1994, 49, 3601-3616.
(10) Salmi, T.; Paatero, E.; Nyholm, P.; Still, M.; Narhi, K.
Kinetics of Melt Polymerization of Maleic and Phthalic Acids with
Propylene Glycol. Chem. Eng. Sci. 1994, 49, 5053-5070.
(11) Lehtonen J.; Salmi, T.; Immonen, K.; Paatero, E.; Nyholm,
P. Kinetic Model for the Homogeneously Catalyzed Polyesterification of Dicarboxylic Acids with Diols. Ind. Eng. Chem. Res. 1996,
35, 3951-3963.
(12) Gordon, M.; Leonis, C. G. Lauric Acid/Pentaerythrityl
Monolaurate: A Model Melt Esterification, Part 1. Kinetics. J.
Chem. Soc., Faraday. Trans. 1975, 71, 161-177.
(13) Durand, D.; Bruneau, C.-M. Study on the gelling of a triol/
diacid/monoacid reaction system with a substitution effect on the
triol. Makromol. Chem. 1977, 178, 3237-3248.
(14) McCarthy, J. L.; Scholtens, B. J. R. Model Esterification
of Neopentyl Glycol and Trimethylol Propane with 1,4-Tert-Butyl
Benzoic Acid. J. Appl. Polym. Sci. 1991, 42, 2223-2231.

4258

Ind. Eng. Chem. Res., Vol. 38, No. 11, 1999

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


Publication Date (Web): October 9, 1999 | doi: 10.1021/ie980633a

(15) Gupta, S. K.; Kumar, A.; Bhargava, A. Molecular weight


distribution and moments for condensation polymerization of
monomers having reactivity different from their homologues.
Polymer 1979, 20, 305-310.
(16) Gupta, S. K.; Kumar, A.; Bhargava, A. Molecular weight
distributions and moments for condensation polymerizations
characterized by two rate constants. Eur. Polym. J. 1979, 15, 557564.
(17) Lopez-Gonzalez, M. M. C.; Callejo Cudero, M. J.; BarralesRienda, J. M. Synthesis of Polyesters as Binders for Deinkable
Inks. 4. Chemical Kinetic Modeling of Copolyesterification between
o-Phthalic Anhydride, Oleic Acid and Neopentyl Glycol in Bulk
at 205 C. J. Phys. Chem. B 1998, 102, 1109-1121.
(18) Robertson, A. M.; Dell, F.; Littlejohn, D.; Brown, M.; Dowle,
C. J.; Goodwin, E. Analysis of an (Isodecyl End-Capped) Propylenediol Adipate Polyester Using Coupled High-Performance
Liquid Chromatography-Fourier Transform Infrared Spectrometry. Anal. Proc. 1993, 30 (6), 268-271.

(19) Hindmarsh, A.C. ODEPACK, A Systemized Collection of


Solvers. In Scientific Computing; Stepleman, R., Ed.; NorthHolland Publishing Co.: Amsterdam, 1983; pp 55-64.
(20) Gear, C. W. Numerical Initial Value Problems in Ordinary
Differential Equations; Prentice Hall: Englewood Cliffs, NJ, 1971.
(21) Marquardt, D. W. An Algorithm for Least Squares Estimation of Non-Linear Parameters. J. Soc. Ind. Appl. Math. 1963, 11,
431-441.
(22) Dennis, J. E.; Gay, D. M.; Welsch, R. E. Algorithm 573
NL2SOL-An Adaptive Nonlinear Least-Squares Algorithm [E4].
ACM Trans. Math. Software 1981, 7, 369-383.

Received for review October 2, 1998


Accepted August 8, 1999
IE980633A

You might also like