Sie sind auf Seite 1von 11

Analysis of Sensitivity of Equilibrium Constant to Reaction

Conditions for Esterication of Fatty Acids with Alcohols


Saeikh Z. Hassan and Madhu Vinjamur*
Department of Chemical Engineering, Indian Institute of Technology Bombay, Powai, Mumbai 400076, India
ABSTRACT: The equilibrium constant (K
eq
) depends on temperature only, but reported measurements for esterication
reactions show that it also depends on reactant concentration and amount of catalyst. In this work, analysis of sensitivity of K
eq
to
errors in equilibrium composition of esterication reactions at dierent reaction conditions is presented. Rigorous error and
Taguchi analyses showed that K
eq
is indeed a function of temperature only and not of molar ratio (MR) of the reactants and
catalyst loading; reported dependency of K
eq
on reactants concentrations and catalyst loadings is due to errors in analysis of
equilibrium composition. Sensitivity of K
eq
to these errors depends strongly on molar ratio (MR), and calculations show that K
eq
is least sensitive to error at 1.5 MR. However, phase separation of the reaction mixture is possible at 1.5 MR which can cause
errors in measured equilibrium composition. K
eq
of esterication reactions should be obtained through kinetic data tting to the
rate model. For sulfuric acid-catalyzed esterication of oleic acid with methanol, a single K
eq
value for each reaction temperature
is obtained through kinetics data tting as 1.53 (0.05), 1.67 (0.04), and 1.91 (0.06) at 313, 323, and 338 K, respectively.
Equilibrium compositions are calculated reasonably well with these K
eq
values.
1. INTRODUCTION
The basic criterion for a reaction to be at equilibrium is that the
change in Gibbs free energy is zero (i.e., G = 0) and thus,
equilibrium constant (K
eq
) is related to standard Gibbs free
energy (G
0
) by the equation G
0
= RT ln K
eq
.
1
For liquid
phase reactions, the thermodynamic K
eq
is dependent on
temperature only and independent of pressure eect except for
high pressures.
1,2
Temperature dependence of K
eq
is expressed
by the vant Ho isochore assuming that the standard change in
enthalpy (H
0
) is constant for a small range of temperatures.
2
For esterication reactions, determination of K
eq
from G
0
has large uncertainty and nonreliability because standard Gibbs
energy of formation (G
f
0
) for each reacting component is high,
and the dierence between the free energies of the products and
reactants (i.e., G
0
) is small. Hence, small errors in measured G
f
0
results in large deviations in K
eq
.
3
Many researchers prefer to
determine K
eq
at dierent temperatures from experimental data
at equilibrium or from tting kinetics data to a reaction rate
model.
Thermodynamic K
eq
, when estimated from thermodynamic
properties of reacting components, depends on temperature only
but the experimental values of K
eq
are reported to depend also on
the composition of the reaction mixture and on the amount of
catalyst.
47
Homogeneous catalysts HCl, HClO
4
, HNO
3
, and
H
2
SO
4
, each aect K
eq
dierently in similar reaction mixtures,
and an increase in K
eq
with catalyst concentration is found.
6
K
eq
at
298 K varied from 1.91 to 4.58 over the range of reaction
compositions studied using acid-ion-exchange resin.
6
The
equilibrium constant of the esterication of ethanol with acetic
acid varied between 2.47 and 4.74, and the dependency on the
molar ratio of the reactants is reported.
8
Othmer and Rao
9
showed that catalyst concentration and reaction temperature
have no signicant inuence on the equilibrium constant of
sulfuric acid-catalyzed esterication of oleic acid with butanol.
There was no signicant inuence of temperature because of the
low endothermicity of the reaction. However, the equilibrium
constant decreased with an increase in molar ratio (alcohol to
fatty acid) as reported by Othmer and Rao
9
and Lee et al.
10
Usually, concentration-based K
eq
(assuming ideal solution) is
considered to depend on many factors, whereas activity-based
K
eq
is considered to depend on temperature only. However,
Pereira et al.
11
reported the variation in activity-based K
eq
of
esterication of lactic acid with ethanol from 3.813 to 4.637 for
the same reaction temperature 323.15 K with dierent
concentrations of initial reactants and catalyst. Pereira et al.
11
attributed this variation in K
eq
to experimental errors as well as
deciencies in the thermodynamic models used to calculate the
activity coecients. Therefore, in order to minimize the error in
equilibrium compositions, Pereira et al.
11
have estimated a
unique K
eq
for each temperature by applying an iterative
procedure and using the experimentally measured equilibrium
compositions. The estimated K
eq
have been applied successfully
in estimation of equilibrium compositions.
From the foregoing, it can be asked if experimentally
determined K
eq
depends on reaction conditions other than
temperature or the reported dependency on the conditions is due
to some other reasons such as error in composition analysis. The
accuracy in analysis is always a matter of concern in experimental
study and use of even high precision analytical tools for
quantitative measurements of reacting agents is prone to errors.
Martin
12
discussed for reactions type R P and R + S P that
show how error in terminal absorbance (using spectropho-
tometer) can cause an inconsistency of calculated K
eq
.
Yalcinyuva et al.
13
and Jong et al.
14
reported the phase splitting
behavior in esterication of myristic acid with isopropyl alcohol
and n-propanol. Since phase splitting causes diculty in
Received: July 15, 2012
Revised: October 21, 2012
Accepted: December 19, 2012
Published: December 19, 2012
Article
pubs.acs.org/IECR
2012 American Chemical Society 1205 dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215
determining the correct overall composition at equilibrium, K
eq
values based on overall composition can be inaccurate.
Experimental equilibrium constants reported by Bart et al.
15
are dierent from the K
eq
calculated by kinetic data tting;
however, these values are reported as independent of catalyst
concentration. Similarly, Omota et al.
16
reported a dierence
between experimental and calculated values for concentration-
based K
eq
. Therefore, experimental errors can cause discrepancy
between data of measured equilibrium constants and data from
kinetics. The eect of experimental errors on K
eq
cannot be
ignored. This eect is assessed in this paper.
In heterogeneous catalysis, properties of solid catalysts and
adsorption anity of reagents at reagent-catalyst interface can
interfere with equilibrium and lead to deviation from true
equilibriumconversion.
17,18
In biocatalysis, the reported depend-
ency of K
eq
on other factors in addition to temperature can be
described dierently from homogeneous and heterogeneous
catalysis.
1921
Bucala et al.,
21
however, attributed inconsistency
in equilibrium constant to experimental and phase equilibrium
calculation errors but the specic properties of biocatalyst are
also responsible. It is reported that higher initial amount of water
improves esterication rates,
22
support of biocatalyst diminishes
the water content from reaction mass,
19
and alcohol can inhibit
the esterication reaction due to its reaction with the active site
triad.
20
Therefore, homogeneous catalysis is a better way to
analyze the true experimental equilibrium position of ester-
ication reactions to determine K
eq
.
In this study, sulfuric acid-catalyzed esterication of oleic acid
with methanol is carried out and the published data for
esterication of fatty acids are also used. The aim of this work
is to understand the sensitivity of equilibrium constant to errors
in equilibriumcomposition analysis and to explore the possibility
of nding a specic condition for a reaction to be carried out
resulting in minimum error of K
eq
.
2. EXPERIMENTAL METHOD
2.1. Reagents. Methanol (GR grade, moisture <0.02%),
water (HPLC grade), sulfuric acid (98% GR), pure oleic acid,
oxalic acid dihydrate (GR grade), n-hexane (HPLC grade), and
phenolphthalein indicator are supplied by Merck India, Ltd.
Rened sunower oil (max. 0.1 wt % FFA) is purchased from
Liberty Oil Mills Ltd., Mumbai, India. Potassium hydroxide
(extra pure AR) is supplied by Sisco Research Laboratories Pvt.
Ltd., Mumbai, India.
2.2. Taguchi Design of Experiments (Taguchi DOE). For
reaction kinetics study, L9 (3
4
) orthogonal array of experiments
(9 experiments, 3 parameters, 3 levels for each parameter) are
conducted for sulfuric acid-catalyzed esterication of 50 wt %
oleic acid (OA) (mixed as free fatty acid in sunower oil) with
methanol. The L9 array is shown in Table 1. First three columns
are lled with reaction parameters and last one is left empty. The
empty column carries two degrees of freedom for residual error
and is used for calculation of the F-ratio in ANOVA. Reaction
parameters are varied at three levels each: catalyst loading, 0.52
wt %; temperature, 313338 K; and molar ratio (MR), 39. MR
is dened as ratio of moles of methanol to moles of oleic acid.
Catalyst loading (wt %) is determined based on the amount of
oleic acid.
2.3. Experiments and Analysis. Oleic acid (OA) is mixed
with rened sunower oil in a 1-L, 4-neck round-bottom ask in
the desired amount. The ask is placed in an oil bath and heated
to the set reaction temperature. Methanol is then added to the
ask and heated again to the set temperature. The calculated
amount of sulfuric acid is added to the reactants mixture in the
ask and the point of admixture of catalyst is considered as the
starting time of the reaction. Through out the reaction, the
mixture is stirred at 500600 rpm. Samples (23 mL each) are
drawn for analysis at dierent intervals and transferred into 25
mL vials. To stop the reaction, 23 mL of n-hexane (cold) is
mixed with a sample, and to separate sulfuric acid, 45 mL of
cold distilled water is then added. Through vigorous mixing
followed by centrifuging for 15 min (4000 rpm) the aqueous and
organic layers are separated. The aqueous layer is removed by a
pipet. Again, 45 mL of water is added to the organic layer and
the mixture is centrifuged to remove traces of sulfuric acid into
water. The organic layer is then transferred to a 10 mL vial which
is kept in a vacuumoven at 70 Cto remove water, methanol, and
n-hexane.
To determine %OA, samples are titrated with standard 0.05 N
potassium hydroxide solution using phenolphthalein as an
indicator.
1
H-NMR technique is used to determine total methyl
esters formed to check the possibility of trans-esterication of
sunower oil as a side reaction. Approximately 0.2 g of each
sample is dissolved in 500 L of deuterated chloroform with a
small portion of TMS (tetramethylsilane) as internal standard
and transferred into a 5-mm diameter tube. The spectra are
recorded using a Mercury plus 300 MHz NMR spectrometer
(Varian, USA), operating at 299.862 MHz. Analysis of variance
(ANOVA), Taguchi analysis, and interpretation of kinetics data
are done with MINITAB 15 software and spreadsheet templates
available on the Web site (www.ee.iitb.ac.in/apte/CV_PRA_
TAGUCHI.htm).
2.4. Calculation of Equilibrium Constant. Equilibrium
compositions of reactants and products are required for
experimental determination of K
eq
. Principally the correct
method for experimental determination of K
eq
is the rigorous
thermodynamic approach with activity coecients. For a
reaction, A + B C + D, the K
eq
=
i
(a
i
)

i
= K

K
C
where a
i
is activity of reacting species i, K

=
i
(
i
)

i
is activity-coecient
equilibrium quotient and K
C
=
i
(C
i
)

i
is concentration
equilibrium quotient.
23,24
Ronnback et al.
24
showed no
signicant improvement in the estimation of K
eq
with the
UNIFAC group distribution method (activity-based approach)
for esterication of acetic acid with methanol; however, the
model t with the concentration-based description is better.
Keurentjes et al.
25
also showed that both approaches,
concentration- and activity-based, predict kinetics of esterica-
Table 1. L9 (3
4
) Orthogonal Array Design of Experiments for
50% Oleic Acid in Oil
expt. no. catalyst loading temperature molar ratio empty
1 1 1 1 1
2 1 2 2 2
3 1 3 3 3
4 2 1 2 3
5 2 2 3 1
6 2 3 1 2
7 3 1 3 2
8 3 2 1 3
9 3 3 2 1
levels
reaction parameters 1 2 3
catalyst loadings (wt%) 0.5 1.0 2.0
temperature (K) 313 323 338
molar ratio (MR) 3 6 9
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1206
tion of tartaric acid with ethanol reasonably, but predictions with
the former perform better. The activity coecients of dierent
reactants may deviate from unity, but this eect can fairly be
compensated as the products of activity coecients (involved in
activity-coecient equilibrium quotient) are close to unity.
25
Ronnback et al.
24
and Keurentjes et al.
25
concluded that kinetics
of acid-catalyzed monoesterication reactions can be studied
with a simple concentration-based rate approach. Thus,
equilibrium constant K
eq
can be treated as equal to K
C
.
3. RESULTS AND DISCUSSION
3.1. Equilibrium Constant (K
eq
). The concentration-based
equilibriumconstants are determined by tting of kinetics data of
esterication of oleic acid with methanol. Each experiment in L9
array is carried out three times and average value of conversion of
oleic acid is taken for kinetic study. Uncertainty in measured
conversion of oleic acid is within 2%. Sulfuric acid is also a
potential catalyst for trans-esterication of oil to methyl esters;
however, acid-catalyzed esterication is many times faster than
trans-esterication.
2628 1
H-NMR is used to calculate total
methyl esters formed by both reactions.
29,30
It is conrmed from
mass balance of oleic acid and total methyl ester that trans-
esterication reaction is negligible. Therefore, trans-esterication
reaction is not considered in kinetics.
The collected kinetics data of esterication of oleic acid in
sunower oil is best tted to reversible second order reaction rate
as shown in Figure 1. Similar kinetics for esterication of fatty
acids with alcohols is also reported by others.
14,3133
Equilibrium
constants are determined by tting of kinetics data of L9
experiments to the analytical solution of second order reversible
kinetics (see caption of Figure 1). Linear least-squares method is
used for data ttings with R
2
> 95%. The equilibriumconstant for
each temperature is found to be 1.53 (0.05) at 313 K, 1.67
(0.04) at 323 K, and 1.91 (0.06) at 338 K, respectively, over
the range of MR and catalyst concentration studied. Figure 2
shows that ln K
eq
vs 1/T gives a straight line which indicates that
K
eq
follows the vant Ho equation d(ln K
eq
)/dT = H/(RT
2
).
Accordingly K
eq
can be expressed as
=

K
T K
30.88exp
941.17
( )
eq
(1)
The enthalpy of reaction is found to be H = +7.82 (0.02) kJ/
mol which is in close agreement with the value reported by
Tesser et al.
34
for the same reaction.
Compositions of reacting agents at equilibrium state can be
used for determining the concentration-based equilibrium
constant as
Figure 1. Reversible second order rate kinetics for esterication of oleic acid with methanol.
=

t
k
K
d[OA]
d
[CH OH][OA]
[H O][MeOA]
3
2
eq
On integration of above equation:
+ + +
+ + +
=

a a X a
a a X a
ka t ln
(1 MR 2 )(1 MR )
(1 MR 2 )(1 MR )
[OA]
2 1 2
2 1 2
2 0
[OA], [CH
3
OH], [MeOA], and [H
2
O] are concentrations of oleic acid, methanol, methyl oleate, and water, respectively. Subscript 0 indicates initial
concentration; k is the rate constant, X is the conversion of OA, K
eq
is the equilibrium constant, MR is the molar ratio, a
1
= 1 (K
eq
)
1
and a
2
= [(1 +
MR)
2
4a
1
MR]
0.5
. For details of experiments see Table 1.
Figure 2. ln(K
eq
) vs 1/T (1/K) plot. The red dashes denote the error
bounds.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1207
= =
=

K
k
k
X
X X
[H O][MeOA]
[CH OH][OA]
(1 )(MR )
eq
forward
backward
2
3
eq
2
eq eq (2)
where X
eq
is the equilibrium conversion of oleic acid (OA).
For esterication of oleic acid with methanol at 346.15 K, an
activity-based equilibrium constant of K
eq
= 2.19 is estimated by
Chen et al.
35
A concentration-based equilibrium constant at
346.15 K is equal to 2.04 when calculated using eq 1. The
dierence between activity-based and concentration-based K
eq
is
less than 7.5%. Chen et al.
35
have applied an iterative algorithmto
estimate the activity-based K
eq
using the UNIQUAC method.
Activity coecients can be calculated by using UNIQUAC to
estimate activity-based K
eq
; these coecients are set to unity in
the concentration-based K
eq
approach. Chen et al.
35
have
considered both the equilibria, chemical reaction equilibrium
and liquidliquid phase (biphasic) equilibrium, in their
algorithm to estimate activity-based K
eq
. The monophasic
concentration-based method is used here in eq 1. Since the
activity-based and concentration-based methods used for the K
eq
estimation are dierent, less than 7.5% dierence between the
estimated K
eq
from both the methods can be considered
reasonable.
In the thermodynamic analysis of equilibrium of sulfuric acid-
catalyzed esterication of oleic acid with methanol, Voll et al.
3
concluded that UNIFAC and modied-Wilson methods are
better in predicting equilibriumthan the ideal solution method (a
concentration-based method with assumption of activity
coecient equal to unity). The results of Voll et al.
3
are shown
in Table 2. It should be noted that Voll et al.
3
used the
experimental data of biodiesel yields obtained after 80 min of
reaction from the study of Lucena et al.
36
as equilibrium data to
validate their estimation methods; Lucena et al.
36
did not report
these yields at equilibrium. In fact, the equilibrium of
esterication reactions usually takes a long time to achieve,
sometimes even days (see from Figure 3). Therefore, 80 min of
reaction duration is insucient for equilibrium analysis.
Equilibrium yields obtained from the long run will be more
than those presented by Voll et al.
3
Voll et al.
3
have reported the
equilibrium yields from an ideal solution assumption, and these
values are found to exceed the experimental values of Lucena et
al.
36
as shown in Table 2.
The equilibrium yields are calculated for similar reaction
conditions (Lucena et al. study
36
) using eqs 1 and 2. Equilibrium
constant at 383.15 K is equal to 2.65 and thus, equilibrium yields
of methyl oleate at 3 and 9 molar ratios (MR) are calculated as
86.70 and 95.69, respectively. Equilibrium yields obtained by
Voll et al.
3
using ideal solution assumption (concentration-based
in Table 2) are in close agreement with our calculated values.
According to reported values of Voll et al.
3
(Table 2), the large
dierence in equilibrium yield calculated from the ideal solution
assumption and activity-based methods indicates a large
deviation from ideal behavior and K

signicantly deviates from


unity. It is already discussed in the preceding paragraph that the
values of experimental biodiesel yield after 80 min of reaction
duration is insucient for validation of equilibrium analysis.
Therefore, their indication seems to be incorrect.
Equilibrium experiments are also carried out for sulfuric acid-
catalyzed esterication of 50 wt % oleic acid (mixed as free fatty
acid in sunower oil) with methanol at 313, 323, and 338 K.
Equilibrium is assumed to be achieved when the oleic acid
conversion rate became less than 0.1% per hour as shown in
Figure 3. Experimental values of X
eq
are compared with
calculated X
eq
, and maximum deviation between them is found
to be 2.81(0.35)%.
Therefore, analysis of equilibrium compositions with concen-
tration-based K
eq
is found to be reasonable. Hence, the
concentration-based approach is valid for kinetics and equili-
brium analysis of homogeneous-catalytic esterication of fatty
acids with alcohols. A similar conclusion is also drawn by
Ronnback et al.
24
and Keurentjes et al.
25
3.2. Eect of Reaction Parameters on K
eq
. 3.2.1. Eect of
Catalyst Loading. An admixture of catalyst accelerates a reaction
and helps attain equilibrium rapidly for reversible reactions.
Reactions proceed even without a catalyst but esterication
reaction rates are low.
37,38
Marchetti and Errazu
26
reported that
the eect of catalyst amount is prominent in the initial phase of
reaction but the equilibrium conversions are the same for
esterication of oleic acid (as free fatty acid) with ethanol using
sulfuric acid.
Goto et al.
7
also studied the eect of catalyst loading on
equilibrium conversion of palmitic acid and their results at 370 K
are shown in Table 3. Theoretical X
eq
is calculated to be 0.9192 at
Table 2. Equilibrium Yields of Biodiesel (Methyl Oleate) in
mol %at 383.15 K, Atmospheric Pressure and Dierent Molar
Ratios (MR) (Voll et al.
3
)
MR
experimental
a
(80 min run;
1 wt % H
2
SO
4
) UNIFAC
modied-
Wilson
concentration-
based
b
3 76.6 0.6 77.84 77.84 88.01
9 88.2 0.3 90.93 89.29 96.22
a
Data is taken from Lucena et al.
36 b
When the activity coecient is
taken as unity.
Figure 3. Conversion of oleic acid (OA) with respect to time.
Experiments are conducted at (i) 313 K, 2.5 wt %catalyst, and 1.67 MR;
(ii) 323 K, 2.5 wt % catalyst, and 1.67 MR; (iii) 338 K, 2.0 wt % catalyst,
and 1.67 MR. MR is mole ratio of methanol to oleic acid.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1208
K
eq
= 2.56 and MR = 5. X
eq
(experimental) values in Table 3 are
nearly the same and within +1% error of calculated X
eq
for all
dierent catalyst loadings. Thus, Goto et al.
7
concluded that the
amount of catalyst has no signicant eect on equilibrium
conversion. However, K
eq
(experimental) values varied by 11%
(2.72 to 3.04) as catalyst loading is changed and this variation is
signicant. At 370 K, 5 MR and 6.3 mol/m
3
catalyst; K
eq
(experimental) is 3.04 which is 19% greater than K
eq
(calculated)
for a small dierence of 1.1% between X
eq
(experimental) and
X
eq
(calculated).
To understand the sensitivity of equilibrium constant to error
in X
eq
, 1% error is introduced in X
eq
(calculated) and K
eq
is
calculated for both upper (+1%) and lower (1%) error bounds
at dierent catalyst loadings. Figure 4 shows that K
eq
(experimental) values at dierent catalyst loadings from the
Goto et al.
7
study are between the upper and lower bounds of
calculated K
eq
. Therefore, K
eq
(experimental) is sensitive to error
in the analysis of X
eq
of palmitic acid and thus, the variation in K
eq
(experimental) is due to error in analysis and not because of
variation in catalyst loading.
3.2.2. Eect of Molar Ratio. The equilibrium constants for
esterication of oleic acid with methanol from kinetics data
tting are found to be 1.53 (0.05) at 313 K, 1.67 (0.04) at 323
K, and 1.91 (0.06) at 313 K, respectively. Equilibrium
conversion (X
eq
) of oleic acid at K
eq
= 1.53 (at 313 K) is
calculated for dierent MR using eq 2. Figure 5 shows a sharp
increase in X
eq
(calculated) from22% to 81% with increase in
MR from 0.25 to 3, whereas, further increase in MR from 3 to 9
led to a small increase in X
eq
(calculated), 81% to 93%.
According to Le Chateliers principle, increase in MR shifts the
reaction in the forward direction and increases the equilibrium
conversion for esterication reactions. To understand the
sensitivity of K
eq
to error in analysis of equilibrium conversion
with respect to MR, 1% and 2% error is induced in X
eq
and
then K
eq
is calculated with respect to MR at dierent
temperatures with eq 2.
The sensitivities of K
eq
to 1% and 2 error in equilibrium
conversion with respect to MR at 313, 323, and 338 K are shown
in Figure 6. Figures 6 panels ac show that the sensitivity of K
eq
to error with respect to MR at dierent temperatures is similar.
Surprisingly, sensitivity of K
eq
to error is nonmonotonous with
respect to MR. K
eq
becomes less sensitive to error as MR
increases from 0.25 to 1; with a further increase in MR, K
eq
becomes more sensitive to error. The sensitivity to error is more
pronounced for MR< 1 than for MR> 1 (Figure 6). K
eq
becomes
more sensitive when an error in analysis increases from1% to 2%
at xed MR. For example, at 0.5 MR and 313 K, the deviation of
K
eq
is +15.79% and +34.76% at +1% and +2% error, respectively
(Figure 6a). Therefore, a small error in X
eq
analysis can cause
signicant deviation in K
eq
from the true value, and the deviation
due to error is also MR dependent. A similar trend is observed by
Sandoval et al.,
19
a slight increase in the experimental conversion
leads to a sharp rise in the calculated value of the equilibrium
constant.
Figure 6 shows that the sensitivity of K
eq
to error with respect
to MR passes through a minimum. The sensitivity of K
eq
to error
is low when MR is between 1 and 3. To nd out the MR at which
the sensitivity of K
eq
to error is least (i.e., the least sensitive MR
position), plots of deviations of K
eq
versus MR (at 313 K) are
shown in Figure 7. The deviations are absolute values of
dierence between K
eq
(from kinetics data t) and K
eq
(at 1%
and 2% error in X
eq
). The lowest sensitivity zone is found
between 1 and 2 MR (see Figure 7), and from the calculation of
deviations the least sensitivity point is found at 1.5 MR in all the
induced error analysis. Sensitivity of K
eq
to positive errors (+1%
and +2%) is higher than the negative errors counterpart (1%
and 2%). Hence, 1.5 MR is the best reaction condition to
determine the K
eq
(experimental) for esterication of oleic acid
with methanol.
Therefore, on the basis of the above sensitivity analysis, if a
certain negative error in X
eq
persists in the composition analysis,
then the K
eq
rst increases till 1.5 MR and then decreases beyond
1.5 MR (see Figure 6). This could be a possible reason for a
decrease in equilibrium constant with an increase in molar ratio
as reported by the Othmer and Rao study
9
and Lee et al.
10
3.2.3. Eect of Reaction Temperature. Equilibrium constant
is a function of temperature; with an increase in temperature, K
eq
increases for endothermic reactions and decreases for exothermic
Table 3. Eect of Catalyst Loading on K
eq
(Experimental)
(Goto et al.
7
) at T = 370 K and MR = 5.
catalyst
loading
(mol/m
3
)
X
eq
(experimental)
K
eq
(experimental)
K
eq
(calculated)
X
eq
a
(calculated)
6.3 0.930 3.04
10.5 0.923 2.72
17.0 0.927 2.89 2.56 0.9192
24.2 0.930 3.04
33.5 0.925 2.80
a
Calculated by using eq 2.
Figure 4. Sensitivity of the equilibriumconstant to 1%error in analysis
of X
eq
of palmitic acid at dierent catalyst loadings and 370 K.
Figure 5. Eect of MR on X
eq
(calculated) of oleic acid at 313 K.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1209
reactions. The equilibrium constant determines the thermody-
namic limit of equilibrium conversion and an increase in K
eq
means that the extent of forward reaction exceeds the backward
reaction thus resulting in higher X
eq
. Figure 8 shows that X
eq
(calculated) increases a little when the temperature of reaction is
raised by 25 K at dierent MRs. For example, at 1 MR, X
eq
(calculated) increases from 55% to 58% as the temperature
increases from 313 to 338 K. This small increase is due to low
endothermicity (H = +7.82 0.02 kJ/mol) of the reaction
under study. However, +1% and +2% error in X
eq
impacts K
eq
signicantly for all MR (Figure 9). That is why in Figure 9, X
eq
is
treated as function of temperature only for all MR to understand
the eect of error in X
eq
on K
eq
. The deviations (or sensitivity) are
shown for positive errors only because K
eq
is more sensitive to
positive error than negative error.
Figure 9 shows that K
eq
becomes more sensitive to error with
an increase in X
eq
for all MR. Comparison of the slope of plots at
dierent MR indicates that K
eq
is more sensitive to error at lower
and higher MR. For MR < 1.5 or MR > 1.5, K
eq
becomes more
sensitive. Slope also increases for all MR when error increases
from 1% to 2%. The least slope and sensitivity of K
eq
to errors
with increase in X
eq
(i.e., due to increase in temperature) is found
at 1.5 MR. Therefore, the least sensitivity of K
eq
to error in X
eq
of
Figure 6. Sensitivity of K
eq
to 1% and 2 error in X
eq
with respect to
MR at (a) 313, (b) 323, and (c) 338 K.
Figure 7. Deviation of K
eq
(at 1% and 2% error in X
eq
analysis) from
K
eq
(calculated) at dierent MR and at 313 K. Absolute deviation of
K
eq
= |K
eq
(at %error in X
eq
) K
eq
(from kinetics data t)|.
Figure 8. X
eq
(calculated) with respect to temperature at dierent MR.
Figure 9. Sensitivity of K
eq
to +1% and +2% errors in X
eq
analysis with
increase in X
eq
. X
eq
is treated as function of temperature only for dierent
MR. Absolute deviation of K
eq
= |K
eq
(at %error in X
eq
) K
eq
(from
kinetics data t)|.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1210
the esterication reaction of oleic acid with methanol for a
temperature range of 313 to 338 K is found at 1.5 MR. Though it
is found that the increase in X
eq
with temperature rise is small for
all MR between 0.25 and 9, the eect of small error in X
eq
on
sensitivity of K
eq
is signicant. For example, at 1.5 MR which is
the least sensitive position, average deviations in K
eq
are +7.56%
and +15.78% for +1% and +2% errors in X
eq
, respectively.
Esterication of fatty acids is usually carried out at high
temperatures (up to 473 K) when solid catalysts are used
(heterogeneous catalysis).
16
If heterogeneous-catalytic reactions
are described by pseudohomogeneous kinetic rate model, K
eq
for
these reactions can be estimated similar to homogeneous-
catalytic reactions. K
eq
is calculated to be 4.22 at 473 K using a
plot of ln K
eq
vs 1/T(or eq 1). To understand the shift in the least
sensitive position of MR with increase in temperature, sensitivity
vs MR at dierent K
eq
(calculated) values are shown in Figure 10.
MR varies from 1 to 2 (lowest sensitive MR zone as shown in
Figure 7), and the sensitivity of K
eq
is calculated for +2% error in
X
eq
. Variation in K
eq
(calculated) from 1.2 to 4.2 covers a large
temperature range, from room temperature to 473 K, for
esterication reactions of oleic acid with methanol. Figure 10
shows that sensitivity of K
eq
to +2% error increases with an
increase in temperature (or K
eq
). The least sensitive position of
MR shifts toward 1 MR with an increase in K
eq
(calculated). For
example, the least sensitive position shifts from1.8 MRto 1.3 MR
as K
eq
(calculated) increases from 1.2 to 4.2. This shift in MR is
not signicant for the large temperature range considered.
Therefore, the least sensitive position of MR is around 1.5 for
esterication of oleic acid with methanol.
The above-discussed sensitivity analysis of K
eq
for identi-
cation of least sensitive position of MR (as shown in Figure 10)
can also be used for nding this position for reactions with
unknown equilibrium constant. Prior to experimental determi-
nation of the equilibrium constant, the sensitivity analysis can be
used to nd out the least sensitive value of MR while using a
range of values for K
eq
within the expected range (from
literature) and assuming +1% (or +2%) error in X
eq
analysis.
Therefore, the analysis gives an approximate least sensitive value
of MR that can be used in experiments to determine the
equilibrium constant. Many esterication reactions of fatty acids
with alcohols are described by reversible second order reaction
kinetics.
7,14,15,25
Performing this analysis for reversible second
order reactions, we nd that the least sensitive value of MR is
between MR 1 and 2 even though K
eq
may vary between 1.1 and
40.
3.3. Comparison of Sensitivity Analysis of Activity- and
Concentration-Based K
eq
. In the literature, activity-based K
eq
is considered to depend on temperature only, whereas
concentration-based K
eq
is considered to be dependent on
many other factors also. However, to understand the sensitivity
of activity-based K
eq
to error in composition analysis, the
sensitivity analysis (by induction of 1% error in X
eq
) is carried
out and compared with that of concentration-based K
eq
. For this
analysis, the data from Ali et al.,
39
the study of esterication of
propionic acid with 1-propanol, are taken.
Sensitivity analysis of equilibrium constants K
C
(concen-
tration-based) and K
a
(activity-based) are given in Table 4. From
this Table, it is clear that both K
C
and K
a
are highly sensitive to
1% error in X
eq
. Sensitivity of K
a
is more than K
C
as the percent
deviation in K
a
is two times more than the percent deviation in
K
C
for both +1% and 1% errors. Hence, if a certain error in X
eq
persists in the composition analysis, then K
a
also deviates fromits
true value which is in agreement with Pereira et al.
11
They found
that, for the same reaction temperature, the activity-based K
eq
are
varied with initial concentrations of reactants and catalyst. The
variation in K
eq
is attributed to experimental errors as well as
deciencies in the thermodynamic models used to calculate the
activity coecients.
3.4. Taguchi Analysis. Equilibriumconversion (X
eq
) of oleic
acid is calculated for all L9 experiments using eq 2 and these
conversions are shown in Table 5. In section 3.2, the eect of
1% and 2% error in X
eq
analysis on sensitivity of K
eq
is
discussed with respect to catalyst loading, MR, and temperature.
In this section, Taguchi analysis is used to quantify and compare
the simultaneous eects of all the reaction parameters on X
eq
(calculated) and K
eq
(at 1% and 2 error in X
eq
). The Taguchi
method is a statistical analysis of a system and utilizes various
orthogonal arrays for the design of experiments.
4042
The
method exploits the variation of all variables simultaneously in
order to understand the impact of each of them on the process.
Taguchi analysis and analysis of variance (ANOVA) are done
with MINITAB 15 software. In Taguchi analysis, SN ratio
(signal-to-noise ratio) is used as an objective function, which
indicates sensitivity of signal (quality characteristic or response
factor) to noise (variation or deviation in reaction parameters).
X
eq
and K
eq
are used as response factors for this analysis;
Figure 10. Change in sensitivity of K
eq
to +2% error in X
eq
at dierent
K
eq
(calculated). Shift in least sensitive position of MR is shown within
12 MR zone.
Table 4. Sensitivity Analysis of Concentration-Based (K
C
) and Activity-Based (K
a
) EquilibriumConstant (K
C
and K
a
Values Taken
from Ali et al.
39
)
% deviation
temperature (K) K
C
K
a
K
C
(at error) K
a
(at error) K
C
K
a
303.15 4.213 33.179 3.85
a
27.868
a
8.616 16.006
323.15 3.919 28.370 4.29
b
33.787
b
9.467 19.093
a
Calculated at 1% error in X
eq
.
b
Calculated at +1% error in X
eq
.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1211
temperature, MR, and catalyst loading are used as reaction
parameters. Larger-is-better SN ratio is applied to X
eq
and K
eq
,
and their results are shown in Figures 11 and 12, respectively.
Larger-is-better means the response factor should be as high as
possible and the objective function to be maximized is expressed
as

=
=
=

n y
SN ratio 10 log (mean square reciprocal quality
characteristic)
10 log
1 1
i
n
i
10
10
1
2
(3)
where y is the response factor and n is the number of
observations. The basis of the selection of larger-is-better SN
ratio can be found in detail elsewhere.
4042
An increase in the SN
ratio (larger-is-better) with a variable implies that the response
factor increases with that variable or vice versa.
Figure 11 shows that X
eq
(as response factor) increases sharply
with MR, less pronounced with temperature and negligibly with
catalyst loading. From ANOVA for the SN ratio, MR contributes
97.52% to X
eq
(calculated), temperature contributes 1.82%, and
catalyst loading contributes only 0.37%. According to the Le
Chateliers principle, an increase in MR and temperature shifts
the reaction in the forward direction for the endothermic
reaction. For the reaction considered here, however, X
eq
is
predominantly aected by MR. Temperature eect is small due
to low endothermicity of the reaction as indicated earlier.
L18 (6
1
x 3
6
) orthogonal array
40
is chosen for Taguchi analysis
of K
eq
to understand the eect of 1% and 2% error in X
eq
analysis on the sensitivity of K
eq
to reaction conditions. L18
means a total of 18 trials and the 6
1
3
6
array means, at
maximum, seven variables can be studied simultaneously with
one column of six levels and rest at six columns of three levels
each. In this study, the rst three columns are lled only with
reaction parameters and the remaining columns are left empty.
The rst column is occupied by MR with six levels varying from
0.25 to 6; the second and third columns are occupied by
temperature and catalyst loading, respectively, with three levels
each. Each empty column carries two degrees of freedom for
residual error and can be used for the calculation of F-ratio in
ANOVA.
Figure 12 shows that catalyst loading does not aect K
eq
for all
the errors in X
eq
analysis. The eect of error on sensitivity of K
eq
to MR is dierent in nature at dierent errors in analysis. For
negative errors, 1% and 2%, Figure 12 panels a and b show
that SN ratio rst increases and then decreases beyond 1.5 MR.
For positive errors, +1% and +2%, Figure 12 panels d and e show
opposite behavior. That is, the SN ratio rst decreases and then
increases beyond 1.5 MR. For zero error in analysis (i.e., at X
eq
),
Figure 12c shows that there is no change in SN ratio with MR
indicating no eect of MR on K
eq
. From ANOVA results, the
contribution of MRto K
eq
decreases from66.25%to 0.00%with a
decrease in error from 2% to zero and this contribution
increases from 0.00% to 77.78% with an increase in error from
zero to +2%. Hence, it can be concluded that the eect of MR
(reactants composition) on experimentally determined K
eq
appears due to error in the X
eq
analysis. From Taguchi analysis,
the introduction of only 1% error in X
eq
causes about 3543%
contribution of MR to K
eq
(Figure 12) which is too large.
Therefore, sensitivity of K
eq
to these errors depends strongly on
molar ratio (MR). The eect of error in X
eq
on sensitivity of K
eq
to temperature is important, and SN ratio increases with
temperature for all the errors in analysis (Figure 12ae). For zero
error in composition analysis (i.e., at X
eq
), Figure 12c and
ANOVA results indicate that temperature is the only parameter
which aects K
eq
. Therefore, K
eq
is a function of temperature
only. Errors in X
eq
analysis (highly possible in experimentation)
lead to the erroneous conclusion that K
eq
also depends on
reactants concentrations and catalyst loading.
3.5. Liquid Phase Behavior. Esterication of fatty acid with
methanol produces fatty acid methyl ester and water. The
reaction mixture could phase separate due to immiscibility of the
fatty acid and water, fatty acid methyl ester (FAME) and water,
and methanol and oil in which the acid is present.
43,44
Liu et al.
43
reported liquidliquid immiscibility of oleic acidmethanol
water (I), FAMEmethanolwater (II), oleic acidmethanol
oil (III) and FAMEmethanoloil (IV) systems in the range
303333 K. Methanololeic acid and oiloleic acid are two
completely miscible systems but methanoloil is partially
miscible. However, high content of oleic acid can increase the
miscibility of methanoloil. A single homogeneous phase is
found in 20 wt % oleic acid in oil for MR up to 10.
43
In this study,
50 wt % oleic acid is present (as free fatty acid) in sunower oil
Table 5. Values of K
eq
and X
eq
(Calculated Using K
eq
) for
Esterication of Oleic Acid with Methanol from L9
Orthogonal Array of Experiments
expt.
no
catalyst loading
(wt %)
temperature
(K) MR K
eq
X
eq
(calculated)
1 0.5 313 3 1.53 80.64
2 0.5 323 6 1.67 90.39
3 0.5 338 9 1.91 94.23
4 1.0 313 6 1.53 89.69
5 1.0 323 9 1.67 93.51
6 1.0 338 3 1.91 83.25
7 2.0 313 9 1.53 92.99
8 2.0 323 3 1.67 81.69
9 2.0 338 6 1.91 91.40
Figure 11. Taguchi analysis of X
eq
(calculated) for L9 experiments.
Larger is better SN ratio is used for analysis.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1212
which is enough for feed composition to form a homogeneous
phase for MR up to 9.
The phase splitting of a product mixture is concluded by Liu et
al.
43
based on ternary phase diagram results, whereas the product
composition of esterication of oleic acid (as free fatty acid) with
methanol is more complex and the product consists of ve
components; oil, oleic acid, FAME, methanol, and water. Phase
splitting causes diculty in determining the correct overall
composition at equilibrium. From the sensitivity analysis, an
equilibrium constant is found to be highly sensitive to error in an
equilibrium composition analysis when MR is varied up to 9.
Computations show that experimentally measured K
eq
at 1.5 MR
is least sensitive to error in composition analysis. Lee et al.
10
reported for esterication of acetic acid (pure) with amyl alcohol
that the product mixtures form a homogeneous liquid phase
when MR is greater than or equal to 3, whereas, a water-rich
phase (the second liquid phase) appears when the feed
composition is close to stoichiometric ratio. Therefore, it is
better to analyze the tendency of the liquid phase segregation at
dierent MR. The liquid composition at the equilibrium state is
calculated for dierent MR at a reaction temperature of 313 K, as
shown in Figure 13.
A comparison between results of Figure 13 and the ternary
phase diagrams of Liu et al.
43
can be done here. Oiloleic acid
Figure 12. Taguchi analysis and ANOVA results of K
eq
for dierent %error in X
eq
using L18 orthogonal array (a) 2% error, (b) 1% error, (c) at X
eq
(zero error), (d) +1% error, and (e) +2% error.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1213
FAME is a completly miscible system, and the addition of water
or methanol causes phase separation. As shown in Figure 13,
below 1 MR, the methanol and water content are too low (1.7
and 2.4 wt % water and methanol, respectively at 1 MR) and this
might not be sucient to cause phase segregation. Beyond 1.5
MR, water content is constant as 2 wt %, and FAME does not
change appreciably. Hence, if the content of water and FAME are
constant and oil, oleic acid, and methanol vary, phase behavior of
this ve-component system can be explained on the basis of the
phase diagram of system-III (oleic acidmethanoloil) fromLiu
et al.
43
Though the presence of FAME and water reduces the
homogeneous region, the relative phase segregation tendency of
our reaction mixture with the respect of MR can be analyzed.
Figure 13 shows that oil and oleic acid decrease while methanol
content increases continuously. Comparing these variations in
the reaction composition and phase diagram of system-III, it can
be concluded that the reaction mixture tends to phase separate
with an increase in MR. Thus, more experimental errors are
possible at high MR. However, the content of water and
methanol even at 1.5 MR (the least sensitivity position) can
cause phase splitting and create diculty in equilibrium
composition analysis.
4. CONCLUSIONS
Sensitivity analysis of K
eq
of esterication reactions showed that
K
eq
is a function of temperature only. Errors in X
eq
(highly
possible in experimentation) lead to the erroneous conclusion
that K
eq
also depends on reactants concentrations and catalyst
loading. Sensitivity of K
eq
to error in equilibrium composition
analysis depends strongly on the molar ratio (MR), and
calculations show that K
eq
is least sensitive to the error in X
eq
at 1.5 MR. The tendency of the reaction mixture toward phase
segregation increases with MR. Thus, experimental errors can be
large at high MR. However, the content of water and methanol
even at 1.5 MR can cause phase splitting and create diculty in
accurate equilibrium composition analysis. Therefore, incon-
sistency is found in experimentally measured K
eq
. The variation
in experimentally measured K
eq
for the same temperature at
dierent initial molar ratios and catalyst loadings is reported by
Pereira et al.
11
and Darlington and Guenther.
6
For sulfuric acidcatalyzed esterication of oleic acid with
methanol, a single K
eq
value for each reaction temperature is
obtained through kinetic data tting, and K
eq
values are found to
be reasonable in equilibrium composition analysis. The
maximum deviation between experimental and calculated X
eq
is
found to be 2.81(0.35)%. Therefore, it is better to obtain K
eq
of
esterication reactions through kinetic data tting to the rate
model.

AUTHOR INFORMATION
Corresponding Author
*Tel.: +91-22-25767218. Fax: +91-22-25726895. E-mail:
madhu@che.iitb.ac.in.
Notes
The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
We thank Professor Prakash R. Apte, Electrical Engineering
Department, Indian Institute of Technology Bombay, for his
helpful discussions and suggestions in the use of the Taguchi
method for this study.

REFERENCES
(1) Hill, C. G. An Introduction to Chemical Engineering Kinetics and
Reactor Design; John Wiley & Sons: NY, 1977.
(2) Smith, J. M.; Van Ness, H. C.; Abbott, M. M. Introduction to
Chemical Engineering Thermodynamics, 6th ed.; McGraw-Hill: Singa-
pore, 2001; pp 458459.
(3) Voll, F. A. P.; Silva, C.; da.; Rossi, C. C. R. S.; Guirardello, R.;
Castilhos, F. de.; Oliveira, J. V.; Cardozo-Filho, L. Thermodynamic
Analysis of Fatty Acid Esterification for Fatty Acid Alkyl Esters
Production. Biomass Bioenerg. 2011, 35, 781788.
(4) Zey, E. G. Esterication. Kirk-Othmer Encyclopedia of Chemical
Techology, 3rd ed.; Wiley: New York, 1978; Vol. 9, p 291.
(5) Sandler, S. I. Chemical, Biochemical, and Engineering Thermody-
namics, 4th ed.; John Wiley & Sons: New Delhi, 2006; pp 730736.
(6) Darlington, A.; Guenther, W. B. EthanolAcetic Acid
Esterification Equilibrium with Acid Ion-Exchange Resin as Catalyst. J.
Chem. Eng. Data 1967, 12, 605606.
(7) Goto, S.; Tagawa, T.; Yusoff, A. Kinetics of the Esterification of
Palmitic Acid with Isobutyl Alcohol. Int. J. Chem. Kinet. 1991, 23, 1726.
(8) Swietoslawski, W. Equilibrium Constant of Esterification. J. Phys.
Chem. 1933, 37, 701707.
(9) Othmer, D. F.; Rao, S. A. n-Butyl Oleate from n-Butyl Alcohol and
Oleic Acid. Ind. Eng. Chem. 1950, 42 (9), 19121919.
(10) Lee, M. J.; Wu, H. T.; Lin, H. Kinetics of Catalytic Esterification of
Acetic Acid and Amyl Alcohol over Dowex. Ind. Eng. Chem. Res. 2000,
39, 40944099.
(11) Pereira, C. S. M.; Pinho, S. P.; Silva, V. M. T. M.; Rodrigues, A. E.
Thermodynamic Equilibrium and Reaction Kinetics for the Ester-
ification of Lactic Acid with Ethanol Catalyzed by Acid Ion-Exchange
Resin. Ind. Eng. Chem. Res. 2008, 47, 14531463.
(12) Martin, R. B. Why Does an Equilibrium Constant Not Appear
Constant? J. Chem. Educ. 1998, 75, 14971498.
(13) Yalcinyuva, T.; Deligoz, H.; Boz, I.; Gurkaynak, M. A. Kinetics and
Mechanism of Myristic Acid and Isopropyl Alcohol Esterification
Reaction with Homogeneous and Heterogeneous Catalysts. Int. J. Chem.
Kinet. 2008, 40, 136144.
(14) Jong, M. C. de.; Feijt, R.; Zondervan, E.; Nijhuis, T. A.; Haan, A. B.
de. Reaction Kinetics of the Esterification of Myristic Acid with
Isopropanol and n-Propanol Using p-Toluene Sulphonic Acid as
Catalyst. Appl. Catal. A: Gen. 2009, 365, 141147.
(15) Bart, H. J.; Reidetschlager, J.; Schatka, K.; Lehmann, A. Kinetics of
Esterification of Levulinic Acid with n-Butanol by Homogeneous
Catalysis. Ind. Eng. Chem. Res. 1994, 33, 2125.
(16) Omota, F.; Dimian, A. C.; Bliek, A. Fatty Acid Esterification by
Reactive Distillation: Part 2Kinetics-Based Design for Sulphated
Zirconia Catalysts. Chem. Eng. Sci. 2003, 58, 31753185.
(17) Grossi, C. V.; Jardim, E. de. O.; Arau jo, M. H. De; Lago, R. M.;
Silva, M. J. da. Sulfonated Polystyrene: A Catalyst with Acid and
Figure 13. Liquid composition (in wt %) of a reaction mixture at
equilibrium state. Equilibrium composition is calculated for the 313 K
reaction temperature at various MR.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1214
Superabsorbent Properties for the Esterification of Fatty Acids. Fuel
2010, 89, 257259.
(18) Akbay, E. O

.; Altiokka, M. R. Kinetics of Esterification of Acetic


Acid with n-Amyl Alcohol in the Presence of Amberlyst-36. Appl. Catal.
A: Gen. 2011, 396, 1419.
(19) Sandoval, G.; Condoret, J. S.; Monsan, P.; Marty, A. Esterification
by Immobilized Lipase in Solvent-Free Media: Kinetic and Thermody-
namic Arguments. Biotechnol. Bioeng. 2002, 78, 313320.
(20) Foresti, M. L.; Errazu, A.; Ferreira, M. L. Effect of Several Reaction
Parameters in the Solvent-Free Ethyl Oleate Synthesis using Candida
Rugosa Lipase Immobilised on Polypropylene. Biochem. Eng. J. 2005, 25,
6977.
(21) Bucala, V.; Foresti, M. L.; Trubiano, G.; Ferreira, M. L.; Briozzo,
M.; Bottini, S. Analysis of Solvent-Free Ethyl Oleate Enzymatic
Synthesis at Equilibrium Conditions. Enzyme Microb. Technol. 2006,
38, 914920.
(22) Foresti, M. L.; Pedernera, M.; Ferreira, M. L.; Bucala, V. Kinetic
Modeling of Enzymatic Ethyl Oleate Synthesis Carried Out in Biphasic
Systems. Appl. Catal. A: Gen. 2008, 334, 6572.
(23) Miao, S.; Shanks, B. H. Mechanism of Acetic Acid Esterification
over Sulfonic Acid-Functionalized Mesoporous Silica. J. Catal. 2011,
279, 136143.
(24) Ronnback, R.; Salmi, T.; Vuori, A.; Haario, H.; Lehtonen, J.;
Sundqvist, A.; Tirronen, E. Development of a Kinetic Model for the
Esterification of Acetic Acid with Methanol in the Presence of a
Homogeneous Acid Catalyst. Chem. Eng. Sci. 1997, 52, 33693381.
(25) Keurentjes, J. T. F.; Janssen, G. H. R.; Gorissen, J. J. The
Esterification of Tartaric Acid with Ethanol: Kinetics and Shifting the
Equilibrium by Means of Pervaporation. Chem. Eng. Sci. 1994, 49,
46814689.
(26) Marchetti, J. M.; Errazu, A. F. Esterification of Free Fatty Acids
Using Sulfuric Acid as Catalyst in the Presence of Triglycerides. Biomass
Bioenerg. 2008, 32, 892895.
(27) Lou, W. Y.; Zong, M. H.; Duan, Z. Q. Efficient Production of
Biodiesel from High Free Fatty Acid-Containing Waste Oils Using
Various Carbohydrate-Derived Solid Acid Catalysts. Bioresour. Technol.
2008, 99, 87528758.
(28) Wang, Y.; Ou, S.; Liu, P.; Xue, F.; Tang, S. Comparison of Two
Different Processes to Synthesize Biodiesel by Waste Cooking Oil. J.
Mol. Catal. A: Chem. 2006, 252, 107112.
(29) Gelbard, G.; Bres, O.; Vargas, R. M.; Vielfaure, F.; Schuchardt, U.
F.
1
H-Nuclear Magnetic Resonance Determination of the Yield of the
Transesterification of Rapeseed Oil with Methanol. J. Am. Oil Chem. Soc.
1995, 72, 12391241.
(30) Satyarthi, J. K.; Srinivas, D.; Ratnasamy, P. Estimation of Free
Fatty Acid Content in Oils, Fats, and Biodiesel by
1
H-NMR
Spectroscopy. Energy Fuel 2009, 23, 22732277.
(31) Altiokka, M. R.; C itak, A. Kinetics Study of Esterification of Acetic
Acid with Isobutanol in the Presence of Amberlite Catalyst. Appl. Catal.
A: Gen. 2003, 239, 141148.
(32) Aafaqi, R.; Mohamed, A. R.; Bhatia, S. Kinetics of Esterification of
Palmitic Acid with Isopropanol using p-Toluene Sulfonic Acid and Zinc
Ethanoate Supported over Silica Gel as Catalysts. J. Chem. Technol.
Biotechnol. 2004, 79, 11271134.
(33) Dhanuka, V. R.; Malshe, V. C.; Chandalia, S. B. Kinetics of the
Liquid Phase Esterification of Carboxylic Acids with Alcohols in the
Presence of Acid Catalysts: Re-interpretation of Published Data. Chem.
Eng. Sci. 1977, 32, 551556.
(34) Tesser, R.; Serio, M. D.; Guida, M.; Nastasi, M.; Santacesaria, E.
Kinetics of Oleic Acid Esterification with Methanol in the Presence of
Triglycerides. Ind. Eng. Chem. Res. 2005, 44, 79787982.
(35) Chen, F.; Sun, H.; Naka, Y.; Kawasaki, J. Reaction and Liquid
Liquid Distribution Equilibria in Oleic Acid/Methanol/Methyl Oleate/
Water System at 73C. J. Chem. Eng. Jpn. 2001, 34, 14791485.
(36) Lucena, I. L; Silva, G. F.; Fernandes, F. A. N. Biodiesel Production
by Esterification of Oleic Acid with Methanol using a Water Adsorption
Apparatus. Ind. Eng. Chem. Res. 2008, 47, 688589.
(37) Sarkar, A.; Ghosh, S. K.; Pramanik, P. Investigation of the
Catalytic Efficiency of a New Mesoporous Catalyst SnO
2
/WO
3
towards
Oleic Acid Esterification. J. Mol. Cat. A: Chem. 2010, 327, 7379.
(38) Cardoso, A. L.; Augusti, R.; Silva, M. J. da. Investigation on the
Esterification of Fatty Acids Catalyzed by the H
3
PW
12
O
40
Hetero-
polyacid. J. Am. Oil Chem. Soc. 2008, 85, 555560.
(39) Ali, S. H.; Tarakmah, A.; Merchant, S. Q.; Al-Sahhaf, T. Synthesis
of Esters: Development of the Rate Expression for the Dowex 50 Wx8-
400 Catalyzed Esterification of Propionic Acid with 1-Propanol. Chem.
Eng. Sci. 2007, 62, 31973217.
(40) Phadke, M. S. Quality Engineering Using Robust Design, 1st ed.;
Pearson Education: Upper Saddle River, NJ, 2008.
(41) Bagachi, T. P. Taguchi Methods Explained: Practical Steps to Robust
Design; Prentice-Hall of India Pvt. Ltd: New Delhi, 1993.
(42) Baker, T. B. Engineering Quality by Design: Interpreting the Taguchi
Approach; Marcel Dekker: New York, 1990.
(43) Liu, Y.; Lu, H.; Liu, C.; Liang, B. Solubility Measurement for the
Reaction Systems in Pre-esterification of High Acid Value Jatropha
Curcas L. Oil. J. Chem. Eng. Data 2009, 54, 14211425.
(44) Zhou, H.; Lu, H.; Liang, B. Solubility of Multicomponent Systems
in the Biodiesel Production by Transesterification of Jatropha Curcas L.
Oil with Methanol. J. Chem. Eng. Data 2006, 51, 11301135.
Industrial & Engineering Chemistry Research Article
dx.doi.org/10.1021/ie301881g | Ind. Eng. Chem. Res. 2013, 52, 12051215 1215

Das könnte Ihnen auch gefallen