Вы находитесь на странице: 1из 9

Article

pubs.acs.org/IECR

Kinetics of Mixed Ethanol/nButanol Esterication of Butyric Acid


with Amberlyst 70 and pToluene Sulfonic Acid
Arati Santhanakrishnan, Abigail Shannon, Lars Peereboom, Carl T. Lira, and Dennis J. Miller*

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Department of Chemical Engineering and Materials Science, Michigan State University, 2527 Engineering Building, East Lansing,
Michigan 48824-1226, United States
ABSTRACT: Esterication of butyric acid with ethanol, n-butanol, and ethanol/n-butanol mixtures was studied using Amberlyst
70 cation-exchange resin and homogeneous p-toluene sulfonic acid as catalysts. The kinetics of individual alcohol esterication
were rst examined in batch reactions at dierent temperatures and catalyst loadings, and then esterication in ethanol/n-butanol
mixtures of varying concentration ratios was characterized. Both nonideal solution and ideal solution kinetic models were
developed. These models accurately predict the esterication of butyric acid by the individual alcohols; a simple additive
combination of the individual alcohol esterication kinetics properly describes mixed alcohol esterication, indicating that the
alcohols do not compete with each other or inhibit esterication when present together. When solution density is included in the
kinetic rate expression to account for the actual concentration of OH groups in ethanol and n-butanol, butyric acid
esterication kinetics with the two alcohols are described by a common rate constant. This rate constant also predicts butyric acid
esterication kinetics with other alcohol combinations, suggesting a generalized esterication rate constant for simple alcohol
esterication of butyric acid over Amberlyst 70 catalyst. These results provide signicant predictive capabilities for simulating
processes such as reactive distillation processes for mixed alcohol esterication.

1. INTRODUCTION
The growing need to reduce dependence on fossil sources for
fuels and chemicals has led to exploration of pathways for their
manufacture from renewable sources. Esters of higher alcohols
(alcohols with more than two carbons) are one such
industrially important class of compounds, and economically
viable processes for making them need to be designed. Blends
of esters are being considered as attractive solvents or additives
to biofuels because of their high energy density and favorable
fuel properties. The most common route for producing esters is
direct esterication of carboxylic acids with alcohols1 using
either homogeneous acid catalysts such as sulfuric acid or ptoluene sulfonic acid (p-TSA) or solid heterogeneous acid
catalysts such as cationic exchange resins.
Mixed alcohol streams from biomass can be obtained in
several ways: from condensation of lower alcohols to higher
alcohols via the Guerbet reaction,24 via FischerTropsch
synthesis of alcohols from synthesis gas,58 or from fusel
alcohols9 produced in ethanol fermentation. Esterication for
biofuel or solvent applications is an attractive use of these
mixed alcohol streams, as it would lead to value-added products
without the need for separation into individual components.
Simultaneous esterication with multiple alcohols is described
in the patent literature.1013
Since esterication reactions are thermodynamically limited,
reactive distillation is a viable option for mixed alcohol
processing.14,15 In simulations of reactive distillation, esterication of a mixture of amyl alcohol and n-butanol with acetic
acid has been examined to compare separation-rst and
reaction-rst schemes.16 Reaction-rst schemes were determined to be more economical. Design of such reactive
distillation schemes for mixed alcohols requires a good
understanding of the kinetics of the reaction system, as it is
generally not known whether the presence of one alcohol
2012 American Chemical Society

accelerates or inhibits the rate of reaction of another, or if


formation of mixed esters leads to transesterication that could
overcomplicate the recovery process.
Recent advances in producing butyric acid via fermentation
of biomass carbohydrates has sparked interest in using butyric
acid as a building block via esterication and other reactions.17
Apart from their potential as biofuel components, ethyl
butyrate and n-butyl butyrate serve as food avoring agents
and green solvents.18 The kinetics of n-butyl butyrate formation
using Dowex19 as an esterication catalyst has been previously
studied and modeled using quasi-homogeneous, EleyRideal,
and LangmuirHinshelwood rate models. The kinetics of ethyl
butyrate formation have not been previously reported.
In this study, the kinetic behavior of butyric acid
esterication with mixed ethanol and n-butanol (Scheme 1) is
investigated using homogeneous (p-TSA) and heterogeneous
(Amberlyst 70 ion-exchange resin) catalysts. Kinetics of
individual ethanol and n-butanol esterication reactions are
rst presented, and the results are then used to predict kinetics
of butyric acid esterication with ethanol/n-butanol mixtures of
varying compositions. The results provide the capability to
predict butyric acid esterication reaction kinetics with any
simple alcohol, a useful result for biorenery process design.

2. MATERIALS AND METHODS


2.1. Materials. Reagent-grade ethanol (200 Proof, Decon
Laboratories, Inc., King of Prussia, PA), n-butanol (99.9%,
Sigma Aldrich Corp., St. Louis, Missouri), n-butyl butyrate
Received:
Revised:
Accepted:
Published:
1845

August 23, 2012


November 15, 2012
December 28, 2012
December 28, 2012
dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

Industrial & Engineering Chemistry Research

Article

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Scheme 1. Esterication of Butyric Acid with Ethanol and n-Butanol

Table 1. Summary of Experiments and Conditions


molar feed ratios
run

temp. (C)

catalyst

ethanol:acid

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31

60
80
100
120
60
80
100
120
60
60
60
80
80
80
80
80
80
60
80
100
120
60
80
100
120
60
60
60
80
80
80

A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
A-70
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA
p-TSA

4.3
2.8
8.7
6.6

0.5
2.6
0.9
1.5
4.5
2.0
0.8
1.9
0.6
2.9
2.8
3.7
3.7

0.4
3.6
5.0
0.5
1.9
3.7

n-butanol:acid

3.7
4.4
4.5
3.3
4.7
2.6
4.2
1.7
3.9
1.9
3.8
2.2
3.0

2.0
3.0
2.8
3.7
1.8
3.3
1.0
2.7
1.9
0.9

(>98%, Sigma Aldrich Corp., St. Louis, MO), ethyl butyrate


(>99%, Sigma Aldrich Corp., St. Louis, MO), water (HPLC
solvent, JT Baker Reagent Chemicals. Phillipsburg, NJ), ptoluene sulfonic acid monohydrate (Spectrum Quality Products, Inc., Gardena, CA), butyric acid (>99%, natural, Sigma
Aldrich Corp., St. Louis, MO), acetonitrile (HPLC grade,
Emanuel Merck Damstadt Chemicals, Philadelphia, PA),
methanol (Sigma Aldrich Corp., St. Louis, MO), and ethyl
octanoate (Sigma Aldrich Corp., St. Louis, MO) were used
without further purication. Gas chromatographic (GC)
analysis of the aforementioned chemicals showed no signicant
presence of impurities except trace amounts of water. Hydranal-

solution density () (kmol m3)

catalyst loading (kg cat./kg soln)

14.9
15.0
15.0
14.9
10.9
10.9
10.9
10.9
14.9
12.8
11.5
12.6
12.9
12.6
11.5
12.6
11.5
14.8
14.9
15.0
15.0
10.9
10.9
10.9
10.9
11.4
13.0
14.4
11.4
12.6
14.0

0.0100
0.0101
0.0075
0.0099
0.0096
0.0101
0.0101
0.0115
0.0109
0.0085
0.0099
0.005
0.0105
0.0077
0.0093
0.0196
0.0097
0.0012
0.0013
0.0012
0.0013
0.0013
0.0013
0.0014
0.0013
0.0012
0.0013
0.0013
0.0013
0.0014
0.0014

coulomat E solution (Riedel-de Haen, Seelze, Germany) was


used in Karl Fischer titrations. Helium (99.995%, AirGas, USA)
was used as carrier gas for GC. The properties of the
heterogeneous cation-exchange resin catalyst Amberlyst 70
(Dow Chemical Co., Midland, MI) are reported in the
literature.20
2.2. Heterogeneous Catalyst Conditioning. As-received
Amberlyst 70 (A-70) was sieved in a series of US-standard
sieves (Dual Manufacturing Co., Chicago, IL), and the 45 +
60 mesh (0.250.35 mm diameter) fraction was used in kinetic
experiments. The resin was washed with methanol multiple
times until the supernatant liquid was colorless and then ltered
1846

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Industrial & Engineering Chemistry Research

Article

study. This is because low loadings of catalyst (0.1 wt % pTSA and 1 wt % A-70), low ratios of alcohol to acid (3:1
6:1), and relatively low temperatures were used in the
experiments.
3.1. Mass Transfer Considerations. Accurate characterization of reaction kinetics requires that experiments be
conducted in the kinetic regime, i.e., at conditions where
external and internal mass transfer resistances do not aect the
reaction rate. Preliminary experiments at varying stirring speeds
showed that conversion rates were unaected above 600 rpm,
implying that the external mass transfer resistances are
negligible at a stirring speed of 800 rpm. To estimate the
inuence of intraparticular mass transfer resistance in the
heterogeneous catalyst reaction, the WeiszPrater criterion was
used.21 The observable modulus was rst calculated by eq 1

to remove excess methanol. The resin was then dried in an


oven at 373 K for 2 days. The dried resin was stored in a sealed
container in a desiccator and removed in required amounts for
reactions. Fresh catalyst was used for each experiment.
To nd the ion-exchange capacity of the catalyst, a known
quantity of dry A-70 was submerged in ethanol for 45 h and
then titrated with NaOH. The average ion-exchange capacity
was found to be 2.35 0.1 equiv of H+/kg, in reasonable
agreement with the value reported by the manufacturer.20
2.3. Kinetic Experiments. Isothermal kinetic experiments
were carried out in 75 mL batch reactors in a Parr 5000
Multireactor system (Parr Instrument Co., Moline, IL). The
reactor system is equipped with temperature and stirring speed
control and with a dip tube on each reactor to collect liquid
samples during reaction. The end of the dip tube is tted with a
2 m stainless steel lter to avoid withdrawing solid catalyst
along with liquid sample.
To begin an experiment, ethanol and n-butanol were weighed
out alone or in predetermined molar ratios and added to the
reactor with a known amount of catalyst (A-70 for
heterogeneous catalysis and p-toluene sulfonic acid for
homogeneous catalysis). The reactor was sealed and heated
until it stabilized at the desired reaction temperature. Stirring
was set to 800 rpm unless otherwise specied. Once the desired
temperature was reached, a specied amount of butyric acid
was added to the reactor through the sample port in a single
shot; the moment of addition was taken as time zero of the
reaction. Total reactant weight came up to approximately 0.040
kg per reaction. Samples of 0.51 mL were withdrawn at
specied time intervals during the kinetic regime (06 h) using
3 mL Luer-Lok tip syringes (Becton Dickson and Co., Franklin
Lakes, NJ) and stored in hermetically sealed vials in a standard
refrigerator at 277 K. Samples to characterize reaction
equilibrium were taken 2448 h after the start of reaction.
2.4. Analysis. The initial water concentration in the
reactants was determined by Karl Fischer titration in an
Aquacounter coulometric titrator AQ-2100 (JM Science Inc.,
Grand Island, NY) and taken into account in calculations.
Analysis of reaction samples was carried out in a Varian 450
gas chromatograph outtted with a thermal conductivity
detector (Varian Medical Systems Inc., Palo Alto, CA).
Reaction samples were diluted 10-fold in acetonitrile containing
11.11 wt % ethyl octanoate as an internal standard. Separation
was done on a 0.53 mm i.d, Aquawax-DA 30 m capillary
column with 1.0 m lm thickness. Helium carrier gas ow rate
was set to 10 mL min1. The following temperature program
was used: initial column temperature 313 K for 2 min, ramp at
10 K min1 to 423 K, ramp at 30 K min1 to 503 K, hold 2 min.
Detector temperature was held at 513 K. Standards of known
composition in the range of interest were prepared and run in
the chromatograph before and after reaction samples to
calibrate the response factor of each component of the reaction.

w =

* ) d p
(robs
CAT

Deff C BA 6

(1)

where r*obs is the observed rate of reaction per weight of


catalyst, CAT is the density of catalyst (assumed to be 1000 kg
m3), CBA is the liquid-phase concentration of butyric acid, dp is
the swelled diameter of catalyst at reaction conditions (eq 2)
d p = d p,dry

Vp,swollen
3

Vp,dry

(2)

where dp is 0.30 mm and Vp,swollen/Vp,dry is determined from a


simple measurement to be 2.0 for both alcohols. The eective
diusivity Deff of butyric acid in alcohol is estimated in eq 3,
where pore tortuosity is assumed to be equal to the inverse of
particle porosity , and DBA is bulk diusivity of butyric acid in
alcohol estimated from the WilkeChang equation22

Deff = DBA = DBA 2

(3)

The Thiele modulus () and eectiveness factor () for butyric


acid esterication are calculated from the observable modulus
w = 2 assuming the reaction is pseudo-rst order in butyric
acid (i.e., excess alcohol) and thus = ((tanh())/()). Values
of were evaluated for butyric acid in each of the alcohols
(Table 2) and found to be 0.93 and 0.96 for ethanol and nbutanol esterication, respectively. These values of indicate
that intraparticular resistances can be neglected.
3.2. Reaction Equilibrium Constants. The equilibrium
constant for a given experiment was estimated by determining
the unchanging composition of the reaction solution after 24
Table 2. Intraparticle Eectiveness Factors for A-70Catalyzed Butyric Acid Esterication
molar feed ratio

3. RESULTS
A list of all experiments conducted along with their conditions
(temperature, initial reactant molar ratio, weight fraction of
catalyst) is given in Table 1. Control experiments to investigate
autocatalysis of the reaction showed negligible rates over the
temperature range studied. Although etherication side
reactions have been observed in other studies involving the
esterication of n-butanol7,13 and ethanol8 at temperatures
above 386 K, no ethers (di-n-butyl ether, diethyl ether, or ethyl
n-butyl ether) were observed in the reaction samples in this
1847

run

temp. (C)

ethanol:acid

1
2
3
4
5
6
7
8

60
80
100
120
60
80
100
120

4.3
2.8
8.7
6.6

n-butanol:acid

3.7
4.4
4.5
3.3

0.99
0.99
0.99
0.98
0.99
0.98
0.98
0.98

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Industrial & Engineering Chemistry Research

Article

Figure 1. vant Ho plots for equilibrium constants of ethanol and n-butanol esterication of butyric acid: () ethanol ideal solution; () ethanol
activity based; () n-butanol ideal solution; () n-butanol activity based. Error bars for each temperature represent one standard deviation.

48 h of reaction. The activity-based equilibrium constant Ka,m


for reaction m is given in eq 4.

is the rate of reaction m per unit volume, and xi is the mole


fraction of component i in the liquid mixture. The parameter
i,m is the ratio of stoichiometric coecients of component i
with respect to the reference component in reaction m.
For this esterication system, eq 5 can be expressed in terms
of total molar density of the liquid phase ( = NT/V) because
the total number of moles is conserved and the reaction volume
is constant during reaction.

NC

K a, m = Kx , mK , m =

(xii)EQ
i ,m

i=1

(4)

Here xi, i, and vi,m represent the equilibrium mole fraction,


activity coecient at equilibrium, and stoichiometric coecient,
respectively, of component i in the reaction mixture. The molefraction-based equilibrium constants (Kx,m) were calculated
directly from the equilibrium composition of the reaction
mixture. The ratio of activity coecients (K,m) accounts for
deviation from ideal behavior; values of activity coecients
were estimated using UNIFAC (universal functional activity
coecient).23 The experimental values of Kx and Ka for butyric
acid esterication with ethanol and n-butanol with A-70 ionexchange resin or p-TSA as catalyst are presented together in
Figure 1. Consistent values of the equilibrium constants were
observed with both catalysts. Scatter in the values of the
equilibrium constant is a result of uncertainty arising from the
necessity measuring the concentration of every species
(including water) participating in the reaction at the
equilibrium state. The enthalpy of reaction (Hr), obtained
from the slope of the trend line in Figure 1, is +18 kJ for
ethanol and +12 kJ for n-butanol esterication of butyric acid,
indicating that the reactions are mildly endothermic. Although
previous studies have reported temperature-independent values
of equilibrium constants in kinetic models,19 here the data in
Figure 1 were used to calculate temperature-dependent
equilibrium constants in the kinetic model described below.
3.3. Kinetic Model Description. In a batch reactor, the
change in number of moles Ni of component i participating in
M reactions can be expressed as

dxi
1
= ( i , mrm)
dt
1

The rate of formation of ester in reaction m, rm, is based on the


law of mass action (e.g., a power law model) with the driving
force for the reversible reaction described in terms of species
activity (ai = xii), reecting the role of chemical potential in
describing reaction equilibrium. For esterication in neat
mixtures of reactants (e.g., no solvent) that constitute a
nonideal liquid phase, the activity of each species is multiplied
by solution density to account for the absolute concentration of
each species (OH and COOH) in the reaction mixture.
This inclusion of solution density is necessary to compare
reaction rates for dierent alcoholacid combinations because
the molar concentration of OH groups decreases as alcohol
molecular weight increases (Table 1). This approach has been
previously dened and used for liquid-phase reactions.24,25
The rate of formation (rm) of butyrate ester (BE) in reaction
m can be expressed in terms of the activity of butyric acid (BA),
alcohol (OH), butyrate ester (BE), and water (W), wCAT, the
catalyst loading in the reaction mixture, , the total molar
density of the reacting uid, k0,m, the pre-exponential factor, and
Ea,m, the activation energy of the rate constant for reaction m.
Ea, m
rm = wCAT2 k 0, m exp
x x
RT BA BA OH OH

dNi
dx
= NT i = ( i , mrm)V
dt
dt
1

(6)

(5)

where NT is the total number of moles in the reactor, M is the


number of reactions in the system, V is the reaction volume, rm
1848

x BEBEx W W

K a, m

(7)

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

Industrial & Engineering Chemistry Research

Article

Table 3. Optimized Kinetic Parameters with 95% Condence Limits and Equilibrium Constants from Experimental Data (T in
Kelvin)
n-butanol

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

ethanol
parameter

ideal

activity

ideal

activity

Ka
Amberlyst 70
ko ((kg solnm3)/ (kg cat.skmol))
Ea (kJ/kmol)
p-toluenesulfonic acid
ko ((kg solnm3)/(kg cat.skmol))
Ea (kJ/kmol)

exp(8.182654/T)

exp(8.852207/T)

exp(7.502339/T)

exp(6.881365/T)

4.1 0.5 103


45 900 3400

6.2 1.6 103


47 300 6100

6.8 1.4 103


46 800 290

11.9 0.5 103


48 300 155

22.9 1.4 103


44 900 182

26.2 1.7 103


45 200 192

21.3 2.6 103


44 900 375

71.8 6.8 103


48 300 300

By inserting the appropriate rate expression (eq 7), eq 6 can be


written for every species in the reaction mixture to give a set of
ordinary dierential equations describing the esterication
system. For the individual ethanolbutyric acid system and the
n-butanolbutyric acid system, these equations were integrated
numerically and reaction parameters determined using the
functions nlint and ode23 from the optimization toolbox in
Matlab 7.12.0. Both the activity-based model described above
and an ideal solution model with all activity coecients i set to
unity were regressed to give rate parameters for the individual
reactions. Optimization was done by minimizing Fmin, where
Fmin2 is dened (eq 8) as the sum of squared dierences
between calculated (xicalcd) and experimental (xiexp) species
mole fractions for all species in all esterication reactions
conducted.
2
Fmin
=

1
n

NC

(xi exp xi calcd)2


samples i = 1

(8)

In eq 8, n is the number of experimental samples withdrawn in


all experiments regressed and Nc is the number of reacting
components in those experiments.
Optimized kinetic parameters with 95% condence limits are
reported in Table 3 for ethanolbutyric acid and n-butanol
butyric acid esterication with A-70 and p-TSA catalysts. The
activation energies in Table 3 are in the expected range of
values (45 10 kJ mol1) for esterication of small aliphatic
carboxylic acids with aliphatic alcohols. The temperaturedependent expression for the thermodynamic equilibrium
constant Ka is reported for each reaction in Table 3.
A comparison of the experimental and predicted mole
fraction proles for individual alcohol esterication is given in
Figure 2 for A-70 heterogeneous catalyst and in Figure 3 for pTSA homogeneous catalyst. It can be seen that the kinetic
model ts the individual alcohol esterication reactions
reasonably well.
The kinetic parameters reported in Table 3 for individual
ethanolbutyric acid and n-butanolbutyric acid esterication
reactions were combined, without adjustment or additional
regression, to model the mixed ethanol/n-butanol esterication
of butyric acid. Comparison of both ideal solution and nonideal
activity-based model predictions of mixed alcohol esterication
with experimental data for several experiments is shown in
Figure 4 for A-70 catalyst and Figure 5 for p-TSA. These plots
clearly show that the simple combined model ts the
experimental data well.
The average deviation of the kinetic model prediction from
experimental data at a given data point is described in eq 9.

Figure 2. Experimental and predicted concentration proles of


individual ethanol and n-butanol esterication in the presence of 1
wt % A-70: (a) run 2 (ethanol, T = 80 C); (b) run 5 (n-butanol, T =
60 C). () Ideal solution model prediction; (---) activity-based
model predictions; () butyric acid; () ethanol; () ethyl butyrate;
() water; () n-butanol; () n-butyl butyrate.

fABS =

1
Ncn

NC

|xi exp xi calcd|


samples i = 1

(9)

For ethanolbutyric acid esterication, the average deviation


per data point over all experiments with A-70 was 0.012 and
1849

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Industrial & Engineering Chemistry Research

Article

Figure 3. Experimental and predicted concentration proles of


individual ethanol and n-butanol esterication in the presence of 0.1
wt % pTSA: (a) run 20 (ethanol, T = 100 C); (b) run 23 (n-butanol,
T = 80 C). () Ideal solution model predictions; (---) activity-based
model predictions; () butyric acid; () ethanol; () ethyl butyrate;
() water; () n-butanol; () n-butyl butyrate.

with p-TSA was 0.015. For n-butanolbutyric acid esterication, the average deviation per data point over all experiments
with A-70 was 0.016 and p-TSA was 0.020. These values
indicate that the experimental data are indeed reasonably
represented by the kinetic models. For the mixed alcohol
esterication, the average deviation of the predicted values from
experimental data was 0.011 for A-70 and 0.013 for p-TSA. The
similarity of these deviations with those from the tted
individual alcohol experiments is verication that the combined
model is a good representation of mixed alcohol esterication.
These deviations thus represent the scatter in the experimental
data collection and analysis procedure, the largest contributor
to which was measurement of water concentration using Karl
Fischer analysis.
The uncertainty in the rate constants at the 95% condence
level (Table 3) are quite reasonable, given the relatively small
number of data points (ranging from 32 to 60 depending on
the data set) used in each regression and the relatively low

Figure 4. Experimental and predicted concentration proles of mixed


alcohol esterication in the presence of 1 wt % A-70: (a) run 13
(ethanol:n-butanol = 1.15:1, T = 80 C); (b) run 11 (ethanol:nbutanol = 0.21:1, T = 60 C); (c) run 9 (ethanol:n-butanol = 1:1, T =
60 C). () Ideal solution model predictions; (---) activity-based
model predictions; () butyric acid; () ethanol; () ethyl butyrate;
() water; () n-butanol; () n-butyl butyrate.
1850

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

Industrial & Engineering Chemistry Research

Article

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

sensitivity of the objective function to values of the kinetic


parameters. This low sensitivity arises from the well-known
compensation eect, wherein dierent combinations of preexponential factors and activation energies can give nearly the
same values of rate constants over a narrow range of
temperatures. The values of actual rate constants at a given
temperature, however, are unique in their t and thus useful for
design or comparison between systems.
The rate of esterication with homogeneous p-TSA catalyst
is substantially higher than with heterogeneous A-70 resin.
Using catalyst loadings, the A-70 acid site density of 2.35 equiv
of H+/kg, and initial species concentrations, the initial turn over
number (TON, kmol BA/kmol H+/h) for ethanol on p-TSA is
70.8 and on A-70 is 16.8. For n-butanol, the initial TON is 40.7
on p-TSA and 11.3 on A-70. The ratio of TON for the two
catalysts is approximately four for both alcohols. We attribute
the lower TON in A-70 not to mass transport eects (see Table
2) but most likely to steric eects associated with limited access
of acid and alcohol to an anchored acid site in the porous solid
catalyst versus the unhindered access acid and alcohol have to
free acid in solution. The similarity in activation energies over
the two catalysts supports this postulate and suggests that the
reaction mechanism is the same on homogeneous p-TSA and
heterogeneous A-70.
The ability of individual alcohol kinetic models to t mixed
alcohol esterication over a wide range of alcohol molar ratios
is strong evidence that the alcohols are not competing or
inhibiting each other during esterication. It is consistent with
the accepted mechanism of acid-catalyzed esterication in
which the carboxylic acid is protonated by the acid site,
followed by the slow, rate-limiting attack of nucleophilic alcohol
on the electron-decient carbonyl carbon. There is evidently a
high enough concentration of protonated butyric acid in the
present case to provide ample opportunity for either ethanol or
n-butanol to attack without interference. The results described
here are in accordance with a similar study examining mixed
acid esterication, which reported that the additive combination
of both esterication reactions also predicted the kinetics of the
mixed acid system well.26
3.4. Generalized Rate Constant for Butyric Acid
Esterication. The absolute rate of butyric acid esterication
(kmol/m3/s) in ethanol is higher than in n-butanol for both
catalysts. This is expected because the absolute concentration
(kmol/m3) of ethanol in solution is higher than n-butanol, and
thus, the hydroxyl group concentration is higher.27 However, in
the rate expression developed here for each catalyst, the rate
constants for ethanol and n-butanol esterication at a given
temperature, determined from the pre-exponential factors and
activation energies for each reaction in Table 3, have the same
values within experimental uncertainty. This result is true for
both heterogeneous A-70 and homogeneous p-TSA catalysts.
The activity kinetic model with solution density included (eq 7)
thus unies esterication of butyric acid with ethanol and nbutanol.
The generalization of esterication rate was extended by
applying the rate expression (eq 7) to the initial rates of a set of
C2C8 alcohol esterication reactions with butyric acid
conducted previously in our laboratory27 over A-70 catalyst.
The resulting rate constant at 60 C for each of the alcohols
examined is given in Table 4. It can be clearly seen that the
values of the rate constant are strikingly similar for all alcohols
examined. This results strongly supports the concept of a

Figure 5. Experimental and predicted concentration proles of mixed


alcohol esterication in the presence of 0.1 wt % p-TSA: (a) run 31
(ethanol:n-butanol = 4.1:1, T = 80 C); (b) run 29 (ethanol:n-butanol
= 0.18:1, T = 80 C); (c) run 30 (ethanol:n-butanol = 1:1, T = 80 C).
() Ideal solution model predictions; (---) activity-based model
predictions; () butyric acid; () ethanol; () ethyl butyrate; ()
water; () n-butanol; () n-butyl butyrate.
1851

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

Industrial & Engineering Chemistry Research

Article

dp = swelled diameter of catalyst in reaction conditions, m


Ea,m = activation energy of the rate constant for reaction m,
kJ kmol1
Fmin = square root of mean of squared absolute residues
fabs = average deviation of predicted vs experimental mole
fraction in kinetic modeling
Ka,m = activity-based equilibrium constant for reaction m
Kx,m = mole-fraction-based equilibrium constant for reaction
m
K,m = ratio of activity coecients of species in reaction m
k0,m = pre-exponential factor of reaction m, kg solnm3(kg
cat.)1 s1 kmol1
NC = number of components in reaction
Ni = number of moles of species i in the reaction mixture,
kmol
NT = total number of moles in the reactor, kmol
rm = rate of reaction m per unit volume of liquid phase, kmol
s1 m3
r*obs = observed rate of reaction per weight of catalyst, mol
s1 kg1
R = ideal gas constant
V = volume of reaction liquid phase, m3
Vp,swollen = bulk volume of swollen catalyst, m3
Vp,dry = dry bulk volume of catalyst, m3
wCAT = catalyst loading in the reaction mixture, kg catalyst
(kg solution)1
xi = mole fraction of component i in the liquid mixture

Table 4. Generalized Rate Constants for Butyric Acid


Esterication Over A-70 for Dierent Alcohols

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

alcohol

i, kmolm3

k 104 (60 C) kg solnm3 (kmolkg cat.s)1

methanol
ethanol
ethanol
propanol
n-butanol
n-butanol
isobutanol
2-EHA
4-heptanol
1-heptanol
1-octanol

23.5
16.4
16.4
12.8
10.4
10.4
10.4
6.2
6.8
6.8
6.1

5.0b
2.3a
3.0b
2.2b
3.0a
3.0b
1.9b
2.8b
2.5b
0.9b
0.9b

This work. bReference 27.

generalized rate constant for esterication of butyric acid over


A-70 catalyst with any alcohol.
As a further generalization of butyric acid esterication, we
calculated the initial reaction rate for n-butyl butyrate formation
over Dowex 50Wx8-400 resin19 at 110 C and a 4:1 BA:nBuOH initial feed ratio. Adjusting the value of the resulting rate
constant to 60 C using an activation energy of 38 200 kJ/
kmol19 and accounting for the dierences in acid site density
between Dowex resin and Amberlyst 70, the rate constant for
their system in a form equivalent to that presented in Table 4 is
1.98 104 m3 solnkg soln/(kg cat.kmols), a remarkably
similar value, given the dierences in the reaction system.

Greek

i = activity coecient of component i in the reaction


mixture
= particle porosity
= intraparticle eectiveness factor
W = WeiszPrater observable modulus
= Thiele modulus
= molar density of reacting liquid phase, kmol m3
CAT = density of catalyst particles, kg m3
i,m = ratio of stoichiometric coecient of component i with
respect to the reference component in reaction m
= pore tortuosity
vi,m = stoichiometric coecient of component i in reaction m

4. CONCLUSIONS
The kinetics of ethanol and n-butanol esterication of butyric
acid in the presence of homogeneous p-toluene sulfonic acid
and heterogeneous ion-exchange resin catalyst Amberlyst 70
have been accurately described by an activity-based kinetic
model with solution density included. Mixed alcohol
esterication kinetics are accurately predicted by an additive
combination of esterication rates of the individual alcohols,
indicating that alcohols do not compete with each other or
inhibit each other in simultaneous esterication reactions. By
including solution density to describe actual concentrations of
OH and COOH in solutions of neat reactants, butyric acid
esterication kinetics with simple C2C8 alcohols are
described by a single rate constant, which serves as a
generalized rate constant for butyric acid esterication over
Amberlyst 70. The additivity of rates and existence of a
generalized rate constant greatly simplies the simulation of
actual biorenery esterication processes with single or multiple
alcohols.

Abbreviations

AUTHOR INFORMATION

Corresponding Author

*Tel.: (517) 353-3928. E-mail: millerd@egr.msu.edu.

Notes

The authors declare no competing nancial interest.

A-70 = Amberlyst 70
BA = butyric acid
BE = butyrate ester
BB = n-butyl butyrate
But = n-butanol
EB = ethyl butyrate
EQ = equilibrium
Eth = ethanol
OH = alcohol
p-TSA = p-toluene sulfonic acid
UNIFAC = universal functional activity coecient
W = water

REFERENCES

(1) Iwasaki, T.; Maegawa, Y.; Ohshima, T.; Mashima, K.


Esterication. Kirk-Othmer Encyclopedia of Chemical Technology; John
Wiley & Sons, Inc.: New York, 2000; Vol. 10 (1), pp 133.
(2) Ueda, W.; Ohshida, T.; Kuwabara, T.; Morikawa, Y.
Condensation of alcohol over solid-base catalyst to form higher
alcohols. Catal. Lett. 1992, 12 (1), 97104.
(3) Carlini, C.; Girolamo, M. D.; Macinai, A.; Marchionna, M.;
Noviello, M.; Galletti, A. M. R.; Sbrana, G. Synthesis of iso-butanol by

NOMENCLATURE AND UNITS


ai = activity of species i in reaction solution
CBA = concentration of butyric acid in liquid phase, kmol
m3
Deff = eective liquid-phase diusivity of butyric acid, m2 s1
DBA = bulk diusivity of butyric acid in alcohol, m2 s1
1852

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

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


Publication Date (Web): January 22, 2013 | doi: 10.1021/ie302267s

Industrial & Engineering Chemistry Research

Article

with Amberlyst 70 ion exchange resin as catalyst. Chem. Eng. J. 2012,


188, 98107.
(27) Venkata, K. S.; Pappu, A. J. Y. Lars Peereboom, Evan Muller,
Carl T. Lira, Dennis J. Miller, Effect of carbon chain length of alcohols
in esterification with butyric acid catalyzed by ion exchange resins.
Bioresour. Technol. 2012, DOI: 10.1016/j.biortech.2012.12.087.

the Guerbet condensation of methanol with n-propanol in the


presence of heterogeneous and homogeneous palladium-based
catalytic systems. J. Mol. Catal A: Chem. 2003, 204205, 721728.
(4) Carlini, C.; Macinai, A.; Raspolli Galletti, A. M.; Sbrana, G.
Selective synthesis of 2-ethyl-1-hexanol from n-butanol through the
Guerbet reaction by using bifunctional catalysts based on copper or
palladium precursors and sodium butoxide. J. Mol. Catal A: Chem.
2004, 212 (12), 6570.
(5) Mazanec, T. J. On the mechanism of higher alcohol formation
over metal oxide catalysts: I. A rationale for branching in the synthesis
of higher alcohols from syngas. J. Catal. 1986, 98 (1), 115125.
(6) Mills, G. A. Status and future opportunities for conversion of
synthesis gas to liquid fuels. Fuel 1994, 73 (8), 12431279.
(7) Xiang, M.; Li, D.; Xiao, H.; Zhang, J.; Qi, H.; Li, W.; Zhong, B.;
Sun, Y. Synthesis of higher alcohols from syngas over Fischer
Tropsch elements modified K/-Mo2C catalysts. Fuel 2008, 87 (45),
599603.
(8) Dry, M. E. The FischerTropsch process: 19502000. Catal.
Today 2002, 71 (34), 227241.
(9) Peralta-Yahya, P. P.; Keasling, J. D. Advanced biofuel production
in microbes. Biotechnol. J. 2010, 5 (2), 147162.
(10) Schulz, E.; Bauer, H.; Merscher, K. D. Coupled production of
two esters. U.S. Patent No. 7115773, 2003.
(11) Thurman, L. R.; Harris, J. B.; McAtee, M. R. Coproduction of
low molecular weight esters of alkanols. European Patent No.
0260572, 1988.
(12) Warner, R. J.; Grajales, G. A. R.; Santiago, F. J. S.; Solorzano, F.
A.; Torres, J. A. Process for the simultaneous coproduction and
purication of ethyl acetate and isopropyl acetate. U.S. Patent No.
6765110, 2000.
(13) van Acker, P. E.; Mathieu, O.; Milner, R. J.; Pacynko, W. F. Ester
co-production. U.S. Patent No. 6093845, 2000.
(14) Harmsen, G. J. Reactive distillation: The front-runner of
industrial process intensification: A full review of commercial
applications, research, scale-up, design and operation. Chem. Eng.
Process.: Process Intensification 2007, 46 (9), 774780.
(15) Stankiewicz, A. Reactive separations for process intensification:
an industrial perspective. Chem. Eng. Process.: Process Intensification
2003, 42 (3), 137144.
(16) Lee, H.-Y.; Yen, L.-T.; Chien, I. L.; Huang, H.-P. Reactive
Distillation for Esterification of an Alcohol Mixture Containing nbutanol and n-Amyl Alcohol. Ind. Eng. Chem. Res. 2009, 48 (15),
71867204.
(17) Zhang, C.; Yang, H.; Yang, F.; Ma, Y. Current Progress on
Butyric Acid Production by Fermentation. Curr. Microbiol. 2009, 59
(6), 656663.
(18) Armstrong, D. W.; Yamazaki, H. Natural flavours production: a
biotechnological approach. Trends Biotechnol. 1986, 4 (10), 264268.
(19) Ju, I. B.; Lim, H.-W.; Jeon, W.; Suh, D. J.; Park, M.-J.; Suh, Y.-W.
Kinetic study of catalytic esterification of butyric acid and n-butanol
over Dowex 50Wx8400. Chem. Eng. J. 2011, 168 (1), 293302.
(20) Amberlyst 70 product data sheet; Rohm and Haas: Philadelphia,
2005.
(21) Weisz, P. B.; Hicks, J. S. The behaviour of porous catalyst
particles in view of internal mass and heat diffusion effects. Chem. Eng.
Sci. 1962, 17 (4), 265275.
(22) Wilke, C. R.; Chang, P. Correlation of diffusion coefficients in
dilute solutions. AIChE J. 1955, 1 (2), 264270.
(23) Hansen, H. K.; Rasmussen, P.; Fredenslund, A.; Schiller, M.;
Gmehling, J. Vapor-liquid equilibria by UNIFAC group contribution.
Ind. Eng. Chem. Res. 1991, 30 (10), 23522355.
(24) Eckert, C. A. Molecular thermodynamics of chemical reactions.
Ind. Eng. Chem. 1967, 59 (9), 2032.
(25) Eckert, C. A.; Hsieh, C. K.; McCabe, J. R. Molecular
thermodynamics for chemical reactor design. AIChE J. 1974, 20 (1),
2036.
(26) Orjuela, A.; Yanez, A. J.; Santhanakrishnan, A.; Lira, C. T.;
Miller, D. J. Kinetics of mixed succinic acid/acetic acid esterification
1853

dx.doi.org/10.1021/ie302267s | Ind. Eng. Chem. Res. 2013, 52, 18451853

Вам также может понравиться