1 s2 0 S0009250907002394 main

background image

Chemical Engineering Science 62 (2007) 3197 – 3217

www.elsevier.com/locate/ces

Synthesis of esters: Development of the rate expression for the Dowex 50

Wx8-400 catalyzed esterification of propionic acid with 1-propanol

Sami H. Ali

, Alia Tarakmah, Sabiha Q. Merchant, Taher Al-Sahhaf

Chemical Engineering Department, Kuwait University, PO Box 5969, Safat 13060, Kuwait

Received 20 July 2006; received in revised form 5 March 2007; accepted 13 March 2007

Available online 19 March 2007

Abstract

The kinetics of the esterification reaction of propionic acid with 1-propanol over the ion-exchange resin Dowex 50Wx8-400 has been studied

in this investigation. Kinetic experiments were conducted using a 1 L Lab-Max system at a stirrer speed of 900 rpm over a temperature range
of 303.15 –333.15 K. The catalyst loading was varied from 10 to 60 g dry cat/L and acid to alcohol molar ratios of 1:1, 1:2, 1:4, 2:1 and 4:1
were employed. The equilibrium constants for this reaction were determined in separate experiments at 303.15, 313.15 and 323.15 K. The
values were equal to 33.18, 30.62 and 28.37, respectively, with a standard enthalpy change of reaction of 6.4 kJ/mol. These values show the
reaction to be mildly exothermic. It was found that both external and internal diffusion limitations did not affect the overall reaction rate. The
conversion of propionic acid increased with increasing temperature and catalyst loading and decreased with increasing initial mole fraction of
acid. The increase in chain length of acid or alcohol or branching had a retarding effect on the conversion. Several kinetic models were tested
to correlate the kinetic data, the pseudo-homogeneous (P-H) model, the Eley–Rideal (E–R) model, the Langmuir–Hinshelwood (L–H) model,
the modified Eley–Rideal (M-E–R) model and the modified Langmuir–Hinshelwood (M-L–H) model. In all models, the activity coefficients
were estimated using UNIFAC to account for the non-ideal thermodynamic behavior of reactants and products. A correction factor for the resin
affinity for water (

) was used in both M-E–R and M-L–H models. The above models predicted the kinetic behavior of the studied system

with an overall error ranging from 1.65% to 13.32%. Water was found to be more strongly adsorbed than other species present in the system.
The M-E–R model between adsorbed 1-propanol and non-adsorbed propionic acid which assumes surface reaction as the rate controlling step,
with

equal to 2, was found to be the best model with the least overall error (1.65%). The activation energy for the esterification was estimated

to be 67.3 kJ/mol by this model.
䉷 2007 Elsevier Ltd. All rights reserved.

Keywords: Batch reactor; Catalysis; Esterification; Kinetics; 1-Propyl propionate; Reaction engineering

1. Introduction

Esterification is an important reaction in the chemical engi-

neering industry. The production of esters plays a major role
in the production of flavors and fragrances, solvents, plasti-
cizers, pesticides and herbicides, medicinal and surface-active
agents. Because of the wide use of esters in the chemical
industry (

Venkateswarlu et al., 1976

), it is believed that the

number of commercial esters exceeds 500, with over 100
manufacturers (

McKetta, 1983

). Esterification of alcohols with

Corresponding author. Fax: +965 4839498.
E-mail address:

samiali@kuc01.kuniv.edu.kw

(S.H. Ali).

0009-2509/$ - see front matter

2007 Elsevier Ltd. All rights reserved.

doi:10.1016/j.ces.2007.03.017

mono-carboxylic acids has been studied extensively by earlier
workers (

Rao et al., 1978; Chang and Yeh, 1984; Dakshina-

murty et al., 1984; Namba et al., 1985; Xu and Chuang, 1996;
Zhang et al., 1998

). These studies involve either homogeneous

or heterogeneous catalyzed esterification reactions.

The most commonly used mineral acid is probably sulfuric

acid because of its effectiveness (

Rao et al., 1979

). But, close

control is required since even modest increases in concentra-
tion or temperature can cause dehydration of alcohols to ethers
or olefins (

McKetta, 1983

). Besides alternative acidic catalysts

like para-toluene sulfonic acid and ion-exchange resins con-
taining sulfonic acids, the tri-chloride and sulfate of aluminum
can also be employed (

Rao et al., 1976, 1979

). The sulfonic

acids have good catalytic activity and generally cause fewer

background image

3198

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

side reactions than sulfuric acid. A somewhat higher molar
concentration of the sulfonic acid may be required in order
to achieve the same reaction rate that can be obtained with a
given quantity of sulfuric acid (

McKetta, 1983

). In industrial

applications, the use of mineral acids as catalysts is limited
because it suffers from several drawbacks. The acid catalyst
recovery is uneconomical (

Sharma et al., 1973

) and high acid

concentration will increase the corrosion rate, which increases
the cost of the operation (

Liu and Tan, 2001

). Furthermore,

dealing with homogeneous catalysts waste is very hard because
they have to be neutralized for product separation (

Bhatia et al.,

1973

), which is considered to be a costly process (

Sharma et al.,

1973

). The use of solid catalyst such as ion-exchange resins

has received great attention. Application of these catalysts has
several advantages; recovery of the catalyst is easily achieved
by filtration (

Dakshinamurty et al., 1984

); continuous operation

in columns is possible (

Nagaraju and Mehboob, 1996

); the

purity of the products is higher compared with homogeneous
catalyst since solid acid catalysts are selective and the formation
of by-products is less significant (

Bhatia et al., 1973

); waste

or disposal problems are eliminated (

Nagaraju and Mehboob,

1996

); isolation of reaction intermediates is possible (

Roy and

Bhatia, 1987

).

El Ewady et al. (1984)

studied the effect of acid structure on

the reaction rate. This study was carried out by the esterifica-
tion reaction of methanol with a homologous series of aliphatic
organic acids in the presence of Amberlite IR-120 as the cat-
alyst. Also,

Awad et al. (1997)

studied the effect of alcohol

structure and molecular weight on the reaction rate constant
for the esterification reaction of propionic acid with different
alcohols over styrene–butadiene–phenol formaldehyde (SBPF)
as a catalyst. Esterification of acetic, propionic and pentanoic
acids with different alcohols using a polymer fiber-supported
sulfonic acid, Smopex-101, as a catalyst was studied by

Lilja

et al. (2002)

. Since there is a lack of information on the influ-

ence of acid and/or alcohol structure on esterification reactions
catalyzed by Dowex 50Wx8-400 it was decided to be one of
the aims of this investigation.

Dakshinamurty et al. (1984)

studied the esterification of

1-propanol with propionic acid using Dowex-50W. The exper-
iments were carried out in a batch reactor. The influence of
different variables on the conversion of the reactants was stud-
ied. An empirical model correlating the specific reaction rate
constant in terms of the studied variables was established. It was
proposed that the surface reaction is the rate controlling step
between adsorbed propionic acid and non-adsorbed 1-propanol.
However,

Dakshinamurty et al. (1984)

did not account for the

system non-ideality or for differences in the adsorption of the
reactants and products. Therefore, the lack of systematic kinetic
studies for the esterification of propionic acid with 1-propanol
with the aim of establishing the influences of reaction parame-
ters on the reaction kinetics along with elucidation of the most
probable reaction mechanism by systematic testing of estab-
lished mathematical models is the main impetus for the current
study.

Several kinetic models have been adopted to describe the ki-

netics of heterogeneous catalytic esterification reactions. The

pseudo-homogeneous (P-H) model is similar to the power-
law model for homogeneous reactions (

Xu and Chuang, 1996;

Pöpken et al., 2000; Lee et al., 2002; Gangadwala et al., 2003

).

The P-H model assumes that surface reaction is the controlling
step and adsorption is negligible for all components. Whenever
the adsorption of the molecules taking part in the reaction oc-
curs, the L–H model is applicable for correlating the kinetic
data (

Bhatia et al., 1973; Lee et al., 2000; Pöpken et al., 2000;

Chiplunkar et al., 2005

). On the other hand, the Eley–Rideal

(E–R) model can be applied when the reaction between one
adsorbed species and one non-adsorbed reactant from the bulk
liquid phase is assumed to occur (

Bart et al., 1996; Liu and

Tan, 2001; Gangadwala et al., 2003

).

Our research group (

Ali and Merchant, 2006

) studied the

esterification reaction of acetic acid with 2-propanol using
different ion-exchange resins (Dowex 50Wx8-400, Amber-
lite IR-120 and Amberlyst 15). Under the studied conditions,
the highest conversion was obtained for the system catalyzed
by Dowex 50Wx8-400 at 4 h and 343 K, 1:1 acid to alcohol
molar ratio and 40 g dry cat/L catalyst loading. It was also
found that the systems catalyzed by gel-type resins (Dowex
50Wx8-400 and Amberlite IR-120) exhibited some similari-
ties in their reaction kinetics. The data were fitted to different
models such as the P-H, the E–R, the L–H and the M-L–H
models.

The purpose of our investigation is to study the reaction of 1-

propanol with propionic acid catalyzed by the cation-exchange
resin Dowex 50Wx8-400. The impact of different variables
such as catalyst loading, temperature and acid to alcohol ra-
tio was investigated. Other factors investigated include the im-
pact of catalyst moisture content on the esterification reaction,
the effect of using sulfuric acid rather than Dowex 50Wx8-
400, and the effect of ion-exchange resin catalyst type and
the impact of the presence of water on the reaction. The reac-
tion of acetic or butyric versus propionic acid with 1-propanol
was compared using Dowex 50 Wx8-400. Also, a comparison
was made between five different alcohols, methanol, ethanol,
1-propanol, 2-propanol and 1-butanol, reacting with propionic
acid in the esterification reaction using Dowex 50 Wx8-400.
On the other hand, the significance of both external and inter-
nal diffusion limitations on the esterification system was stud-
ied. Several kinetic models were tested to correlate the kinetic
data, namely the P-H model, the E–R model, the L–H model,
the modified Eley–Rideal (M-E–R) model and the modified
Langmuir–Hinshelwood (M-L–H) model. The non-ideality of
the system behavior was accounted for by universal functional
activity coefficient (UNIFAC).

2. Theory

2.1. Reaction and reaction mechanism

Esters can be formed by the reaction of a carboxylic acid

with an alcohol forming the ester and water molecules. This es-
terification (reversible) reaction, also called the intermolecular
dehydration reaction, is a very important and a common type
of reaction in the chemical industry. The general esterification

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3199

reaction is shown below:

acid

+ alcohol ↔ ester + water.

(1)

The catalytic esterification reaction of a propionic acid and 1-
propanol to form 1-propyl propionate and water is given by

CH

3

CH

2

COOH

+ CH

3

CH

2

CH

2

OH

k

f

←→

k

b

CH

3

CH

2

COOCH

2

CH

2

CH

3

+ H

2

O.

(2)

The reaction mechanism for the formation of esters from car-
boxylic acids and alcohols in the presence of acidic catalyst
was presented by

Lilja et al. (2002)

. The reaction is initiated

by the transfer of a proton from the catalyst to the carboxylic
acid. The proton becomes attached to one of the lone pairs on
the oxygen which is double-bonded to the carbon. The trans-
fer of the proton to the oxygen gives it a positive charge. This
results in a fair amount of positive charge on the carbon atom.
Then, the positive charge on the carbon atom is attacked by the
hydroxyl group of the alcohol molecule. After that, a molecule
of water is lost from the ion. Finally, the catalyst is recovered
by the transfer of proton from the ion to the catalyst surface.
This mechanism is represented by the following scheme:

R

O

OH

H- CAT

+

R

OH

OH

+

CAT

R

OH

OH

+

HO-R'

R

OH

O

+

H

O

H

+

CAT

R'

R

OH

O

R'

R

O

O

R'

H- CAT

+

The donation of a proton is commonly assumed to be a fast
step, while the nucleophilic substitution is usually assumed to
be slow followed by fast steps resulting in the formation of
ester and water and the recovery of the catalyst.

2.2. Diffusion

To have a purely kinetic study, it is necessary to eliminate

both external and internal diffusion limitations. For our case
study, where the reaction of propionic acid with 1-propanol over
solid catalyst was carried out in a batch reactor, the external
mass transfer resistance to the esterification reaction is directly
related to the stirrer speed. The effect of external diffusion lim-
itation on the esterification reaction rate was studied by earlier
workers (

Krishnaiah and Rao, 1984; Yadav and Kulkarni, 2000;

Yadav and Thathagar, 2002; Ali and Merchant, 2006

). To study

the external diffusion effect on the reaction rate, different stirrer

speeds should be applied to the reaction system. If the produc-
tion of the ester is independent of stirrer speed, this indicates
that external diffusion is not the rate controlling step. Thus, to
ensure that the reaction rate is not influenced by external dif-
fusion, the experiments should be run at a high enough stirrer
speed. In general, external diffusion controls the overall rate
in catalytic processes if the viscosity of the reactant mixture is
very high or the stirrer speed is very low (

Othmer, 1994

).

The effect of internal diffusion on the rate of the reaction cat-

alyzed by a solid catalyst (ion-exchange resin) is dependent on
many parameters such as catalyst composition, particle size, re-
action medium and temperature. The effect of internal diffusion
on the catalytic reaction can be studied by screening catalyst
into different particle sizes or by calculating certain dimension-
less parameters such as the well-known Weisz–Prater criterion.
Earlier workers studied the significance of internal diffusion on
the esterification reactions (

Bhatia et al., 1973; Krishnaiah and

Rao, 1984; Bart et al., 1996; Gangadwala et al., 2003; Pääkkönen
and Krause, 2003; Ali and Merchant, 2006

). Some of these

studies tested the reaction for internal diffusion limitation by
using different particle sizes (

Bhatia et al., 1973; Gangadwala

et al., 2003; Pääkkönen and Krause, 2003

), while other studies

used certain criterion for such a purpose (

Krishnaiah and Rao,

1984; Bart et al., 1996; Ali and Merchant, 2006

). The mea-

sured values of the rate of the reaction (

r

A(obs)

) are used to

calculate the Weisz–Prater criterion in order to determine the
possibility of internal diffusion limiting the reaction. The di-
mensionless Weisz–Prater parameter (C

W P

) can be defined as

follows (

Fogler, 1992

):

C

W P

= −

r

A(obs)

c

R

2

c

D

e

C

li

,

(3)

where

r

A(obs)

is the rate of the reaction at a given time in

mol/g of catalyst/s,

c

is the catalyst density in g/cm

3

, R

c

is

the effective radius of the catalyst is the ratio of catalyst pellet
volume to catalyst pellet external surface area in cm, D

e

is the

background image

3200

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

effective diffusivity in cm

2

/s and C

li

is the limiting reactant

concentration in the mixture at a given time in mol/cm

3

.

The definition of D

e

is given below:

D

e

=

2

v

D

lm

,

(4)

where

v

is the void fraction and D

lm

is the diffusivity of limiting

reactant in the mixture in (cm

2

/s).

The multi-component diffusivity was calculated using Perkin

and Geankoplis method (

Poling et al., 2001

) as shown below:

D

lm

0.8

m

=

n

i=1

i=l

x

i

D

li

0.8

i

,

(5)

where x

i

is the mole fraction of the component i, D

li

is the

binary diffusivity of limiting reactant in component i (calculated
by Tyn and Calus method), and

i

and

m

are the viscosity of

component i and the mixture in cp, respectively. The viscosity of
the pure components and the mixture is obtained from HYSYS
program version 3.1.

All the terms in Eq. (3) are either measured or known. If

C

W P

>1 there are no internal diffusion limitations and no con-

centration gradient exists within the pellet, but if C

W P

?1 in-

ternal diffusion limits the reaction (

Fogler, 1992

).

2.3. Kinetic models

Different models were tested in our investigation. These

models are based on different assumptions regarding the re-
action mechanism and the rate controlling step. The precise
reaction mechanism needs to be defined. For instance, for a
situation where adsorption is the rate controlling step, there are
possibly three different reaction mechanisms. Adsorbed pro-
pionic acid is reacting with 1-propanol in the fluid, adsorbed
1-propanol is reacting with propionic acid in the fluid or ad-
sorbed propionic acid is reacting with adsorbed 1-propanol.
Also, the rate controlling step can be the surface reaction step
or the step involving desorption of any of the products. Com-
bining the reaction mechanism with the rate controlling step in
our case (for the reaction of propionic acid and 1-propanol) 13
models are possible as shown in

Table 1

. By taking into ac-

count the strong water affinity for the cation-exchange resins
(

Lee et al., 2000, 2002; Gangadwala et al., 2003; Ali and

Merchant, 2006

) a correction term (

) was added to the activity

term for water in these equations (with

values equal to 1, 2

and 3) resulting in 39 different rate expressions. The values of
the correction term for the affinity of resin for water (

) were

chosen according to previous workers results (

Lee et al., 2000,

2002; Gangadwala et al., 2003; Ali and Merchant, 2006

). In

addition to these 39 possibilities, the P-H model was tested,
resulting in a total of 40 different rate expressions (see Table
1). Therefore, the main differences between these models are
the assumed reaction mechanism and the rate controlling step
in addition to the value of the correction term for water affin-
ity. The P-H is the simplest model, where the adsorption and
desorption of reactants and products are neglected. The second
type of models, the L–H, can describe the kinetic data based

on the assumption that both reactants are adsorbed. However,
if a model derived on the basis of the assumption that one of
the adsorbed reactants is reacting with another in the fluid, then
it can be described by the E–R model. Because the affinity of
the resin for water is strong, the activity of water in the rate
expression is raised to a power

, which can take values of 2

and 3 (

> 1) in the M-L–H and the M-E–R models.

In

Table 1

, M

cat

is the mass of the catalyst in gram, n is the

total number of moles, a

i

is the activity of species i(a

i

=

i

x

i

),

i

is the activity coefficient of species i, x

i

is the mole fraction

of species i, k

f

is the forward reaction rate constant in mol/g

catalyst/s, K

a

is the overall reaction equilibrium constant, K

i

is the adsorption constant for species i, and

is the correction

term for strong water affinity for the cation-exchange resins.

The general expression for the overall reaction equilibrium

constant is given by

K

a

= K

s

K

acid

K

alc

K

ester

K

water

,

(6)

where K

s

is the experimentally measured (surface reaction)

equilibrium constant, K

acid

, K

alc

, K

ester

and K

water

are the ad-

sorption equilibrium constants for acid, alcohol, ester and wa-
ter, respectively. For the P-H model, K

a

= K

s

, while for the

L–H model, K

a

is as in Eq. (6). However, in the case of the

E–R model, only two species (one reactant and one product)
are being adsorbed. So, there are four possibilities of K

a

ex-

pressions with one adsorption equilibrium constant term of one
of the reactants in the numerator of Eq. (6) and another one for
one of the products in the denominator.

2.4. Estimation of activity coefficients

Some correlations for determining the activity coefficient

uses the contributions of interactions between functional
groups, such as –OH,

&O, –CH

3

, etc. rather than the interac-

tion between the molecules in the mixture. The determination
of the activity coefficient can be done without the need for ac-
tual data, by assembling the pure component from individual
groups and assessing the contributions of their interactions.
One of these methods is the UNIFAC (universal functional
activity coefficient) model. UNIFAC was used to account for
system non-ideality. We have found that UNIFAC accurately
predicted the system non-linearity (

Ali and Merchant, 2006

).

The activity coefficient consists of two parts, combinatorial
and residual parts, and it is expressed as

ln

i

= ln

C

i

+ ln

R

i

.

(7)

The combinatorial part of the activity coefficients (

Hansen

et al., 1992

) is given by

ln

C

i

= ln

i

x

i

+

z
2

q

i

ln

i

i

+ l

i

i

x

i

j

x

j

l

j

,

(8)

where the coordination number z

= 10, x

i

is the mole fraction

of species i and l

i

is given by

l

i

=

z
2

(r

i

q

i

) (r

i

− 1).

(9)

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3201

Table 1
Rate expressions for different rate controlling mechanisms

Adsorption status of reactants

Limiting step

Rate expression

a

Non

Surface reaction

r

i

= M

cat

k

f

(a

acid

a

alc

1

K

a

a

ester

a

water

)

Adsorbed propionic acid reacting
with 1-propanol in the fluid

Surface reaction

r

i

=

M

cat

k

f

K

acid

(a

acid

a

alc

(a

ester

a

water

/K

a

))

(1 + K

acid

a

acid

+ K

water

a

water

)

Adsorption of acid

r

i

=

M

cat

k

acid

(a

acid

(a

ester

a

water

/(K

a

a

alc

)))

(1 +

K

acid

K

a

(a

ester

a

water

/a

alc

) + K

water

a

water

)

Desorption of ester

r

i

=

M

cat

K

a

k

ester

(a

alc

a

acid

/a

water

(a

ester

/K

a

))

(1 + K

acid

a

acid

+ K

a

K

ester

(a

alc

a

acid

/a

water

))

Desorption of water

r

i

=

M

cat

K

a

k

water

(a

alc

a

acid

/a

ester

(a

water

/K

a

))

(1 + K

acid

a

acid

+ K

a

K

water

(a

alc

a

acid

/a

ester

))

Adsorbed 1-propanol reacting with
propionic acid in the fluid

Surface reaction

r

i

=

M

cat

k

f

K

alc

(a

acid

a

alc

(a

ester

a

water

/K

a

))

(1 + K

alc

a

alc

+ K

water

a

water

)

Adsorption of alcohol

r

i

=

M

cat

k

alc

(a

alc

(a

ester

a

water

/(K

a

a

acid

)))

(1 +

K

alc

K

a

(a

ester

a

water

/a

acid

) + K

water

a

water

)

Desorption of ester

r

i

=

M

cat

K

a

k

ester

(a

alc

a

acid

/a

water

(a

ester

/K

a

))

(1 + K

alc

a

alc

+ K

a

K

ester

(a

alc

a

acid

/a

water

))

Desorption of water

r

i

=

M

cat

K

a

k

water

(a

alc

a

acid

/a

ester

(a

water

/K

a

))

(1 + K

alc

a

alc

+ K

a

K

water

(a

alc

a

acid

/a

ester

))

Adsorbed propionic acid reacting
with adsorbed 1-propanol

Surface reaction

r

i

=

M

cat

k

f

K

acid

K

alc

(a

acid

a

alc

(a

ester

a

water

/K

a

))

(1 + K

acid

a

acid

+ K

alc

a

alc

+ K

ester

a

ester

+ K

water

a

water

)

2

Adsorption of acid

r

i

=

M

cat

k

acid

(a

acid

(a

ester

a

water

/(K

a

a

alc

)))

(1 +

K

acid

K

a

(a

ester

a

water

/a

alc

) + K

alc

a

alc

+ K

ester

a

ester

+ K

water

a

water

)

Adsorption of alcohol

r

i

=

M

cat

k

alc

(a

alc

(a

ester

a

water

/(K

a

a

acid

)))

(1 +

K

alc

K

a

(a

ester

a

water

/a

acid

) + K

acid

a

acid

+ K

ester

a

ester

+ K

water

a

water

)

Desorption of ester

r

i

=

M

cat

K

a

k

ester

(a

alc

a

acid

/a

water

(a

ester

/K

a

))

(1 + K

acid

a

acid

+ K

alc

a

alc

+ K

a

K

ester

(a

alc

a

acid

/a

water

) + K

water

a

water

)

Desorption of water

r

i

=

M

cat

K

a

k

water

(a

alc

a

acid

/a

ester

(a

water

/K

a

))

(1 + K

acid

a

acid

+ K

alc

a

alc

+ K

ester

a

ester

+ K

a

K

water

(a

alc

a

acid

/a

ester

))

a

a

water

= (x

water

water

)

.

Surface area fraction and volume fraction are as follows:

i

=

q

i

x

i

j

q

j

x

j

,

(10)

i

=

r

i

x

i

j

r

j

x

j

.

(11)

Parameters r

i

and q

i

are calculated as the sum of the individ-

ual group volume and surface area parameters R

k

and Q

k

as

follows:

r

i

=

k

v

(i)

k

R

k

,

(12)

q

i

=

k

v

(i)

k

Q

k

,

(13)

where v

(i)

k

is the number of k groups in molecule i.

The residual part of the activity coefficient (

Hansen et al.,

1992

) is given by

ln

R

i

=

k

v

(i)

k

(ln

k

− ln

(i)

k

),

(14)

where

k

is the group residual activity coefficient in the mixture

and

(i)

k

is the residual activity coefficient of group k in a

reference solution containing only molecules of types i (pure
component). The

k

or

(i)

k

is calculated by

ln

k

(or

(i)

k

)

= Q

k

1

− ln

m

m

mk

m

m

km

n

n

nm

.

(15)

Here

m

is the area fraction of group m and it is calculated

according to the following equation:

m

=

Q

m

X

m

n

Q

n

X

n

,

(16)

background image

3202

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

where Q

m

is a group parameter and X

m

is the mole fraction of

group m in the mixture:

X

m

=

i

v

(i)

m

x

i

i

n

v

i

n

x

i

.

(17)

The group interaction coefficient

mn

(

Hansen et al., 1992

) is

given by

mn

= exp

a

mn

+ b

mn

T

T

.

(18)

3. Experimental

3.1. Catalysts

Three ion-exchange catalysts were used; Dowex 50Wx8-400

(Aldrich catalog # 217514), Amberlyst 15 (Aldrich catalog #
216380) and Amberlite IR-120 (Aldrich catalog # 216534).
These catalysts were used after they have been vacuum-dried
at a temperature of 343 K for 48 h. Drying at much higher
temperatures (temperature > 373 K) could lead to the loss of
active sulfonic acid sites of the catalysts. The properties of the
ion-exchange resins are shown in

Table 2

. The homogeneous

catalyst used, sulfuric acid (product # UN1830), was supplied
by BDH and had a purity of greater than 99% and a specific
gravity of 1.835.

3.2. Chemicals

1-Propanol with a purity of 99.5% was supplied by Schar-

lau (catalog # AI0437). 2-Propanol with a purity of 99.5%
was supplied by Fluka (catalog # 59300). Methanol analytical
grade (99.8% pure) was supplied by AJAX chemicals (UN #
1230). Ethanol reagent grade with 99.8% purity was supplied
by Scharlau (catalog # ET0016). 1-Butanol reagent grade of
99.5% purity was supplied by Fluka (catalog # 19420). Propi-
onic acid having a product # 81912 and Butyric acid with a cat-
alog # 19215 both have a purity of 99% and were supplied by
Fluka, while acetic acid with a purity of 99.8% was supplied by
Riedel-de Haen (product # 27225). The purity of all alcohols

Table 2
Properties

a

of the cation exchange resins used

Catalyst

Dowex 50Wx8-400

Amberlite IR-120

Amberlyst 15

Manufacturer

Dow Chemical Co.

Rohm & Haas

Rohm & Haas

Supplier

Aldrich

Aldrich

Aldrich

Catalog #

217514

216534

216380

Polymer type

Gel-type

Gel-type

Macro reticular

Matrix type

Styrene–divinyl benzene (DVB)

Styrene–divinyl benzene (DVB)

Styrene–divinyl benzene (DVB)

Functional group

Sulfonic acid

Sulfonic acid

Sulfonic acid

Standard ionic form

H

+

H

+

H

+

Total exchange capacity (meq/mL)

1.7

1.9

1.8

Cross-linking (% DVB)

8

8

20

Moisture content (% mass)

54

45

< 1.6

Particle size range (mm)

0.04–0.07

0.30–1.20

0.30–1.20

a

As reported by the manufacturer.

and acids was checked by gas chromatographic analysis and
found to be comparable to the listed values mentioned above.

For titration purposes, the alkali used was a standard solution

of sodium hydroxide (NaOH) with 0.1024 N in water. This so-
lution was supplied by Aldrich (product # 31,948-1). The con-
centration of the alkali solution was confirmed by back titrating
with a freshly prepared solution of potassium hydrogen phtha-
late of known concentration. Potassium hydrogen phthalate was
supplied by Aldrich (catalog # 17992-2) and had purity greater
than 99.9%.

3.3. Kinetic runs

All the kinetic runs were carried out in a Lab-Max reactor

system. The equipment used in this experiment (

Fig. 1

) con-

sisted mainly of a 1 L glass batch reactor system. The reac-
tor was continuously stirred using a four blade glass impeller
driven by an electrical motor. The shell of the reactor vessel
was filled with oil to either heat or cool the reaction mixture. A
temperature probe was inserted into the reactor to measure the
mixture’s temperature with an accuracy of

±0.1 K. The reactor

temperature was automatically controlled. The Lab-Max had a
stirrer speed ranging from 0 to 1500 rpm. A measured amount
of acid and catalyst (liquid or vacuum-dried solid catalyst) was
added to the reactor, and the temperature of the reactor was
raised to the required reaction temperature. A certain amount
of alcohol was separately preheated to the reaction temperature
in a heating bath. The preheated alcohol was added to the re-
actor and the time of initiation of the reaction was noted (zero
time). A sample of the reaction mixture was immediately with-
drawn, filtered and quenched to stop the reaction. The sample
was titrated with a standard NaOH solution. The reproducibil-
ity of the titration results was found to be within

±1.5%. This

was done by repeating the titration procedure three times. Other
samples were withdrawn and titrated at different time intervals
and the reaction was followed for 240 min.

3.4. Equilibrium runs

The equilibrium runs were carried out in 80 cm

3

glass cells

having outer glass jackets facilitating water flow to control the

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3203

Fig. 1. Experimental setup of the Lab-Max apparatus.

temperature from a circulating water bath. The glass stoppers
for the cells were fitted with smaller diameter stoppers at the
top to facilitate the removal of the reaction sample by means
of a thin graduated pipette. This involved removing only the
small upper stopper. Equimolar amounts of propionic acid and
1-propanol (around 0.2 mol of each) along with 2% by volume
of concentrated sulfuric acid were allowed to react in the cells
until equilibrium was reached (as evidenced by the absence of
any change in the amount of alkali required to neutralize the
acid present in 0.1 ml of the withdrawn sample). The reaction
volume was not allowed to exceed half the volume of the cells.
The reaction temperatures studied were 303.15, 313.15 and
323.15 K.

4. Results and discussion

The reaction conditions were systematically altered to study

the effect of rpm, nature of catalyst, catalyst loading, temper-
ature, acid to alcohol molar ratio, presence of water, changing
acid and alcohol types on the reaction kinetics. In addition,
three runs were carried out using sulfuric acid as the catalyst to
determine the reaction equilibrium. The details of the reaction
conditions of these runs are given in

Table 3

.

4.1. External and internal diffusion significance

The esterification reaction in this investigation is a

liquid–solid catalytic reaction where different processes are
taking place: external and internal diffusions, adsorption of at
least one of the reactants, surface reaction and desorption of
products followed by back diffusion of the products into the liq-
uid bulk. To study the kinetics of the esterification reaction, the
effect of external and internal diffusion limitations should be
eliminated. The external mass transfer resistance is affected by
the speed of agitation in our case as has been stated previously.

Therefore, five experiments were carried out at different stirrer
speeds (100, 200, 600, 900 and 1000 rpm). It was found that
the conversion of acid was independent of stirrer speed except
at 100 rpm as shown in

Fig. 2

, because at 100 rpm the catalyst

was not well distributed inside the reactor. This means that for
the setup used in this study (stirrer of 5 cm diameter

× 1 cm

bore diameter and reactor of 10 cm internal diameter), the
external diffusion limitation is negligible at stirrer speeds of
200 rpm and above. Therefore, a stirrer speed of 900 rpm was
maintained during all experiments to ensure that the measured
reaction rate was free from external diffusion effects.

Xu and

Chuang (1996)

studied the methyl acetate synthesis using Am-

berlyst 15 and they found that there was no external diffusion
effect for stirrer speed ranging from 160 to 760 rpm. In the
study undertaken by

Xu and Chuang (1996)

, low rpm values

were found to have a negligible effect on reaction rate because
of the low viscosity of the studied system (reaction of methanol
with dilute acetic acid solution). Furthermore,

Pöpken et al.

(2000)

studied the effect of stirrer speed (ranged from 100

to 560 rpm) for methyl acetate synthesis using Amberlyst 15,
and found that at stirrer speeds of 170 rpm and above, external
diffusion did not control the overall rate in the ion-exchange
resin catalyzed process. This result is in line with the work
of

Xu and Chuang (1996)

.

Roy and Bhatia (1987)

found that

a stirrer speed of 500 rpm was sufficient to eliminate external
diffusion limitations for the esterification reaction of benzyl
alcohol with acetic acid with Amberlyst 15 catalyst.

Ali and

Merchant (2006)

found that in the reaction of 2-propanol with

acetic acid, external mass transfer limitation was negligible
at a stirrer speed above or equal to 500 rpm (the stirrer speed
studied range was from 50 to 900 rpm). A higher range of
stirrer speeds (700–1200) rpm was studied by

Krishnaiah and

Rao (1984)

for the esterification of 1-propanol with acetic acid

over Dowex-50W. The absence of a stirrer speed effect was
observed at 1100 and 1200 rpm, which is much higher than

background image

3204

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

Table 3
Summary of experimental conditions

Run #

Temperature (K)

Acid

to

alcohol

rounded molar
ratio

Catalyst
loading
(g dry cat/L)

Stirrer speed
(rpm)

Acid

Alcohol

Water

Catalyst

1

323.15

1:1

40

100

Propionic

1-Propanol

Dowex 50Wx8-400

2

323.15

1:1

40

200

Propionic

1-Propanol

Dowex 50Wx8-400

3

323.15

1:1

40

600

Propionic

1-Propanol

Dowex 50Wx8-400

4

323.15

1:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

5

323.15

1:1

40

1000

Propionic

1-Propanol

Dowex 50Wx8-400

6

323.15

1:1

60

900

Propionic

1-Propanol

Amberlyst 15

7

323.15

1:1

60

900

Propionic

1-Propanol

Amberlite IR-120

8

323.15

1:1

60

900

Propionic

1-Propanol

Dowex 50Wx8-400

9

323.15

1:1

10

900

Propionic

1-Propanol

Dowex 50Wx8-400

10

323.15

1:1

20

900

Propionic

1-Propanol

Dowex 50Wx8-400

11

323.15

1:1

60

900

Propionic

1-Propanol

Dowex 50Wx8-400

12

323.15

1:1

40

900

Propionic

1-Propanol

Wet Dowex 50Wx8-400

13

323.15

1:1

2% Volume

900

Propionic

1-Propanol

H

2

SO

4

14

303.15

1:1

2% Volume

900

Propionic

1-Propanol

H

2

SO

4

15

303.15

1:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

16

313.15

1:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

17

333.15

1:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

18

323.15

2:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

19

323.15

1:2

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

20

323.15

1:4

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

21

323.15

4:1

40

900

Propionic

1-Propanol

Dowex 50Wx8-400

22

323.15

1:1

40

900

Propionic

1-Propanol

Water added

Dowex 50Wx8-400

23

323.15

1:1

40

900

Acetic

1-Propanol

Dowex 50Wx8-400

24

323.15

1:1

40

900

Butyric

1-Propanol

Dowex 50Wx8-400

25

323.15

1:1

40

900

Propionic

Methanol

Dowex 50Wx8-400

26

323.15

1:1

40

900

Propionic

Ethanol

Dowex 50Wx8-400

27

323.15

1:1

40

900

Propionic

1-Butanol

Dowex 50Wx8-400

28

323.15

1:1

40

900

Propionic

2-Propanol

Dowex 50Wx8-400

29

303.15

1:1

2% Volume

Equilibrium runs

Propionic

1-Propanol

H

2

SO

4

30

313.15

1:1

2% Volume

Propionic

1-Propanol

H

2

SO

4

31

323.15

1:1

2% Volume

Propionic

1-Propanol

H

2

SO

4

0

0.1

0.2

0.3

0.4

0.5

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

100 rpm
200 rpm
600 rpm
900 rpm
1000 rpm

Fig. 2. Effect of stirrer speed (rpm) on the conversion of propionic acid at
323 K, 1:1 propionic acid to 1-propanol molar ratio and 40 g dry cat/L catalyst
loading of Dowex 50Wx8-400.

that found by

Xu and Chuang (1996)

and

Pöpken et al. (2000)

.

Therefore, the significance of external mass transfer limitation
which is directly related to stirrer speed in batch systems de-
pends on several factors such as the viscosity of the system,
reactions conditions, type of species used and the presence of
diluents, in addition to the type and properties of catalyst used.

To investigate the internal diffusion effect on the reaction

rate, analysis based on the Weisz–Prater criterion was under-
taken. Data from 10 experiments (4, 10, 11 and 15–21) were
fitted to the Weisz–Prater equation where Dowex 50Wx8-400
was the catalyst. The Weisz–Prater criterion was determined
for the initial stages (1800 s) of runs 4, 10, 11 and 15–21 (these
runs were used for modeling purpose). The results are shown
in

Table 4

. It was found that values of the internal diffusion

parameter are significantly less than one (C

W P

>1). These

results indicate that internal diffusion does not limit the reac-
tion of propionic acid with 1-propanol over Dowex 50Wx8-
400 for the reaction conditions implemented in this study. The
C

W P

parameter values ranges between 3.63E

−4 to 1.22E −3.

These values are of the same order as those obtained by

Bart

et al. (1996)

and

Krishnaiah and Rao (1984)

, which varied

from 1.3E

− 4 to 1.0E − 2 and from 6.0E − 4 to 1.7E − 3, re-

spectively. Both studies were undertaken for the esterification

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3205

Table 4
Significance of internal diffusion for different runs

Run #

Experimental parameters

a

C

li

at 30 min (mol/cm

3

)

r

A

(obs)

at 30 min (mol/g

cat

/s)

D

e

(cm

2

/s)

C

W P

4

40/323.15/1

0.0055

8.78E

−06

1.22E

−06

8.73E

−04

10

20/323.15/1

0.0058

1.02E

−05

1.16E

−06

1.02E

−03

11

60/323.15/1

0.0052

7.84E

−06

1.31E

−06

7.73E

−04

15

40/303.15/1

0.0060

2.35E

−06

7.23E

−07

3.63E

−04

16

40/313.15/1

0.0059

3.80E

−06

9.21E

−07

4.70E

−04

17

40/333.15/1

0.0048

1.48E

−05

1.69E

−06

1.22E

−03

18

40/323.15/2

0.0077

7.51E

−06

9.00E

−07

7.29E

−04

19

40/323.15/0.5

0.0036

6.22E

−06

1.44E

−06

8.03E

−04

20

40/323.15/0.25

0.0021

4.01E

−06

1.58E

−06

7.97E

−04

21

40/323.15/4

0.0096

6.39E

−06

5.83E

−07

7.71E

−04

a

First number: catalyst loading in g dry cat/L; second number: temperature in K; third number: acid to alcohol molar ratio.

Table 5
Equilibrium mole fractions, conversions, activity coefficients and constants

T (K)

Acetic acid

1-Propanol

Propyl propionate

Water

Equilibrium
constant K

s

Equilibrium

Equilibrium

Activity

Equilibrium

Equilibrium

Activity

Equilibrium

Activity

Equilibrium

Activity

mole

conversion

coefficient

mole

conversion

coefficient

mole

coefficient

mole

coefficient

fraction

fraction

fraction

fraction

303.15

0.1638

0.6724

0.6842

0.1638

0.6724

1.0698

0.3362

1.9273

0.3362

2.9911

33.1789

313.15

0.1654

0.6692

0.7101

0.1654

0.6692

1.0505

0.3346

1.9083

0.3346

2.9248

30.6202

323.15

0.1678

0.6644

0.7239

0.1678

0.6644

1.0376

0.3322

1.8815

0.3322

2.8897

28.3703

of 1-propanol with acetic acid;

Bart et al. (1996)

used Dowex

monospheres, while

Krishnaiah and Rao (1984)

used Dowex-

50W as the catalyst. In addition, in a previous work, our group
found C

W P

to range from 9.3E

− 5 to 5.0E − 4 for the esterifi-

cation of 2-propanol with acetic acid over Dowex 50Wx8-400
(

Ali and Merchant, 2006

).

4.2. Effect of temperature on reaction equilibrium

Three experimental runs were carried out at 303.15, 313.15

and 323.15 K to obtain the equilibrium constant values. The
alcohol to acid molar ratio of 1:1 and sulfuric acid with a
concentration of 2% by volume were used. The experiments
were undertaken to determine the equilibrium mole fractions of
propionic acid, 1-propanol, propyl propionate and water. The
results for the equilibrium runs are shown in

Table 5

.

The following equation was used to determine the equilib-

rium constant K

s

:

K

s

=

(x

ester

)

eq

(x

water

)

eq

(x

acid

)

eq

(x

alc

)

eq

ester

water

acid

alc

,

(19)

where (x

i

)

eq

is the equilibrium mole fraction of component i

and

i

is the activity coefficient of component i determined

by the UNIFAC model. The values of the activity coefficients
are reported in

Table 5

. From

Table 5

it is obvious that the

activity coefficients of 1-propanol, 1-propyl propionate and
water decreased and the activity coefficient of propionic acid
increased as the reaction temperature increased. Increasing the

3.32

3.34

3.36

3.38

3.4

3.42

3.44

3.46

3.48

3.5

3.52

0.00305

0.0031

0.00315

0.0032

0.00325

0.0033

0.00335

1/T (1/K)

ln Ks

Experimental Run

Linear fit

Fig. 3. The natural logarithm of equilibrium constant versus the reciprocal
of the absolute temperature.

temperature from 303.15 to 323.15 K, the equilibrium conver-
sion of the reactants decreased from 0.6724 to 0.6644.

The plot of (ln K

s

) versus (1/T ) with a linear regression of

R

2

fit

= 0.99 (

Fig. 3

) gives the following equation:

K

s

= 2.6445 Exp(6376/RT ).

(20)

background image

3206

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

This equation is consistent with the Van’t Hoff equation for
the effect of temperature on the chemical reaction equilibrium
constant:

ln

K

S

K

SR

= −

H

0

R

R

1

T

1

T

R

.

(21)

Accordingly, the standard enthalpy change of reaction

H

0

R

is

−6.4 kJ/mol. Therefore, the reaction of propionic acid with

1-propanol is exothermic, and the equilibrium constant has a
weak dependence on temperature. The same trend was ob-
tained by

Song et al. (1998)

for the methyl acetate synthe-

sis with

H

0

R

equal to

−6.5 kJ/mol. The equilibrium constant

was found to have values of 30.2, 27.4 and 24.0 at 313.15,
318.15 and 323.15 K, respectively. Also, in a previous work
(

Ali and Merchant, 2006

) we have found that

H

0

R

was equal

to

−5.4 kJ/mol for the esterification reaction between acetic

acid and 2-propanol, and the equilibrium constant to be 29.4,
26.0 and 23.0 at 303, 323 and 343 K, respectively. These re-
sults indicate that the reaction is mildly exothermic. Different
results were obtained by

Bart et al. (1996)

for the esterification

of acetic acid with 1-propanol and

Liu and Tan (2001)

for the

esterification of propionic acid with 1-butanol. They found that
the equilibrium constant increased with increasing temperature,
which indicates that these esterification reactions are endother-
mic.

Bart et al. (1996)

found that the equilibrium constants at

303.15, 323.15 and 343.15 K were 10.3, 21.7 and 32.3, respec-
tively, while

Liu and Tan (2001)

found that the corresponding

values at 363, 373 and 383 K are 29.1, 29.4 and 30.1, respec-
tively. It has to be mentioned that the equilibrium constant is
found to be independent of temperature for many esterification
reactions, because the heat of reaction for these reactions were
almost equal to zero or quite small in value (

McKetta, 1983

).

In this study the performance of several models (in which

different components are assumed to be adsorbed to different
extents on the catalyst surface) were compared. The P-H model,
which is the simplest model used, assumes reaction homogene-
ity and its rate equation calls for a homogeneous reaction equi-
librium constant (K

s

) as given by Eq. (20). In all other models

the reaction is heterogeneously catalyzed by Dowex 50Wx8-
400. Therefore, the equilibrium constant (K

a

) is a combination

of adsorption equilibrium constants (K

i

) for different species

i in addition to the surface reaction equilibrium constant (K

s

)

as expressed in Eq. (6). Since the functional group in the solid
catalyst (Dowex 50Wx8-400) used is sulfonic acid, it was
considered useful to obtain values of the surface reaction
equilibrium constant (K

s

) from a reaction system catalyzed

homogeneously by sulfuric acid at the same catalyst load-
ing. The homogeneous reaction equilibrium constant (K

s

)

was maintained for all other models in order to be able to
have a clear direct comparison of different species adsorption
equilibrium constants (K

i

) calculated by these models.

4.3. Effect of reaction parameters

By eliminating both external and internal diffusion limita-

tions, the esterification reaction is purely kinetically controlled.

0

0.1

0.2

0.3

0.4

0.5

0.6

0

3600

7200

10800

14400

18000

Time (sec)

Convesion of propionic acid

Amberlyst 15

Amberlite IR-120

Dowex 50Wx8-400

Fig. 4. Effect of catalyst type on the conversion of propionic acid at 323 K,
900 rpm, 1:1 propionic acid to 1-propanol molar ratio and 60 g dry cat/L
catalyst loading.

4.3.1. Effect of type of ion-exchange resin catalyst

Ion-exchange resins have been found to be suitable catalysts

for esterification reactions (

Sharma et al., 1973; Dakshina-

murty et al., 1984; Xu and Chuang, 1996; Yadav and Thathagar,
2002

). In recent studies, sulfuric acid is being replaced

by ion-exchange resins due to environmental regulations.
Moreover, using ion-exchange resins has several advantages as
discussed previously. In this investigation, three types of ion-
exchange resins (Dowex 50Wx8-400, Amberlite IR-120 and
Amberlyst 15) were chosen for this purpose. Among these
catalysts, Dowex 50Wx8-400 was found to be the best, yield-
ing the highest conversion of propionic acid as shown in

Fig.

4

. The performance of different ion-exchange resin catalysts

for this kind of reactions is actually attributed to the type of
reactants and products involved, in addition to the structure
and characterization of the ion-exchange resins.

4.3.2. Effect of using sulfuric acid

(H

2

SO

4

) compared with

Dowex 50Wx8-400 as a catalyst

A comparison of the effect of using sulfuric acid rather than

Dowex 50Wx8-400 on the conversion of propionic acid as a
function of time was made at two different temperatures (303.15
and 323.15 K) with identical experimental conditions (

Fig. 5

).

In these experiments, equivalent catalyst loadings were used
for both Dowex (40 g dry cat/L) and sulfuric acid (almost 2%
of total liquid volume so as to equal 40 g dry cat/L). Higher
conversion of propionic acid over a period of time of 4 h was
obtained with sulfuric acid rather than Dowex 50Wx8-400 at
the two temperatures studied. Since sulfuric acid is in the same
phase as the reactants (homogeneous liquid phase), the avail-
ability of free protons in this liquid–liquid catalytic reaction

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3207

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0

3600

7200

10800

14400

18000

Time (sec)

Conversionof propionic acid

H2SO4 at 323.15 K

Dowex 50Wx8-400 at323.15 K
H2SO4 at 303.15 K

Dowex 50Wx8-400 at303.15 K

Fig. 5. Effect of H

2

SO

4

compared with Dowex 50Wx8-400 on the conversion

of propionic acid at different temperatures, 900 rpm, 1:1 acid to alcohol molar
ratio and 40 g dry cat/L catalyst loading.

mixture results in having faster rates of reaction compared to
the case where ion-exchange resins are used (liquid–solid re-
action mixture). However, taking into consideration the advan-
tages of ion-exchange resin, Dowex 50Wx8-400 is the preferred
catalyst.

Roy and Bhatia (1987)

compared the behavior of the

ion-exchange resin Amberlyst 15 with the behavior of sulfuric
acid as catalyst for the conversion of acetic acid as a function
of time. The concentration of catalyst and the reaction condi-
tions were maintained the same. They (

Roy and Bhatia, 1987

)

found that sulfuric acid resulted in a higher conversion than
Amberlyst 15. However, by taking into consideration that us-
ing sulfuric acid as a catalyst suffers from several drawbacks
in industrial application, Amberlyst 15 was found to be more
effective as a catalyst for the synthesis of benzyl alcohol with
acetic acid. The effect of using a homogeneous catalyst versus
heterogeneous catalyst for the esterification of propionic acid
with 1-butanol was investigated by

Liu and Tan (2001)

. Sulfuric

acid was compared with Amberlyst 35 among other solid cat-
alysts at the same catalyst concentration of 1 wt% in solution.
They found that using a concentration of 1 wt% Amberlyst 35
in solution gave less conversion than a concentration of 1 wt%
of sulfuric acid in solution. However, Amberlyst 35 was se-
lected as the best choice for this synthesis because it withstands
higher temperatures than the other tested solid catalysts.

4.3.3. Effect of catalyst loading

The effect of catalyst loading of Dowex 50Wx8-400 on the

conversion of propionic acid was investigated by setting the
catalyst loadings at 10, 20, 40 and 60 g dry cat/L. It is observed
that the conversion of propionic acid increases proportionally
with catalyst loading (see

Fig. 6

). As expected, increasing the

0

0.1

0.2

0.3

0.4

0.5

0.6

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionicacid

10 g dry cat/L

20 g dry cat/L
40 g dry cat/L
60 g dry cat/L

Fig. 6. Effect of Dowex 50Wx8-400 loading on the conversion of propionic
acid at 323 K, 900 rpm and 1:1 propionic acid to 1-propanol molar ratio.

catalyst loading means more available active sites for this reac-
tion, which results in higher reaction rate. A similar trend was
obtained by earlier workers (

Rao et al., 1979; Xu and Chuang,

1996; Bart et al., 1996; Yadav and Kulkarni, 2000

). This re-

sult agrees with the model rate expression where the term M

cat

is explicitly added. This clearly indicates that M

cat

is directly

proportional to the reaction rate.

4.3.4. Effect of the catalyst moisture content (wet versus dry
Dowex 50Wx8-400)

To study the effect of moisture content of the ion-exchange

resin on the reaction, two experiments were conducted at the
same reaction conditions for a time period of 4 h as shown
in

Fig. 7

. The first experiment was undertaken using a wet

catalyst (stock sample) where the weight of water was taken
into account, while for the second experiment the catalyst was
vacuum-dried at a temperature of 343 K for 48 h. The two ex-
periments were carried out using the same dry mass of cata-
lyst. As shown in

Fig. 7

, the conversion of acid was higher

with dry catalyst compared with wet catalyst. Earlier work-
ers used different dried ion-exchange resins (Dowex 50W-x8
and x2, Amberlyst 15 and 35, Amberlite IR-120) for studying
the kinetics of the esterification synthesis (

Sharma et al., 1973;

Krishnaiah and Rao, 1984; Pöpken et al., 2000; Gangadwala
et al., 2003; Ali and Merchant, 2006

).

Sharma et al. (1973)

oven-dried Dowex 50w-x8 at 353.15–358.15 K for the ester-
ification of ethanol with propionic acid. Amberlyst 15 was
vacuum-dried at 363.15 K for 48 h by

Pöpken et al. (2000)

for the methyl acetate synthesis. Furthermore,

Krishnaiah and

background image

3208

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

0

0.1

0.2

0.3

0.4

0.5

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

Wet Dowex 50Wx8-400

Dry Dowex 50Wx8-400

Fig. 7. Effect of moisture content (wet versus dry Dowex) on the conversion
of propionic acid at 323 K, 900 rpm, 1:1 propionic acid to 1-propanol molar
ratio and 40 g dry cat/L catalyst loading.

Rao (1984)

, who studied the esterification of 1-propanol with

acetic acid, preheated Dowex 50W-x2 and x8 in an air oven at
338.15 K for 20 min before use. They found that the moisture
content of the solid ion-exchange resins had a retarding effect
on the conversion of acid. Moreover, a recent study on homo-
geneously catalyzed esterification (

Liu et al., 2006

) has estab-

lished that besides promoting ester hydrolysis, the presence of
water significantly decreases the activity of the catalytic pro-
tons. The presence of water can be expected to have a simi-
lar effect for the studied heterogeneous systems also, since the
catalyst used is acidic in nature. Therefore, it is important to
dry the catalyst before running an experiment since the pres-
ence of water can promote ester hydrolysis and adversely effect
catalytic activity. Both these factors can cause considerable de-
crease in the acid conversions. The role of water in heteroge-
neous esterification is further discussed in the section “Effect
of adding water to the reaction mixture”.

4.3.5. Effect of temperature

The effect of temperature was studied by varying the reaction

temperature from 303.15 to 333.15 K at constant conditions of
acid to alcohol molar ratio, catalyst concentration and stirrer
speed.

Fig. 8

shows that increasing the temperature brings more

collisions and therefore more successful collisions. These suc-
cessful collisions have sufficient energy (activation energy) to
break the bonds and form products and thus result in higher
values of conversion of propionic acid, which agrees with pre-
vious esterification studies (

Venkateswarlu et al., 1976; Rao

et al., 1976; Sai, 1988; Lee et al., 1999; Awad et al., 1997;

0

0.1

0.2

0.3

0.4

0.5

0.6

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

303.15 K
313.15 K
323.15 K
333.15 K

Fig. 8. Effect of temperature on the conversion of propionic acid at 900 rpm,
1:1 propionic acid to 1-propanol molar ratio and 40 g dry cat/L catalyst loading
of Dowex 50Wx8-400.

Ali and Merchant, 2006

). The temperature 323.15 K was se-

lected as the standard temperature for subsequent runs because
it is high enough to yield a high conversion, while gas-bubble
formation is insignificant.

4.3.6. Effect of acid to alcohol molar ratio

The effect of acid to alcohol molar ratio was investigated

by varying the acid to alcohol molar ratio (ranging from 1:4
to 4:1). From

Fig. 9

it is obvious that increasing the amount

of alcohol initially increases the conversion of propionic acid.
After 4 h of reaction time the conversion of propionic acid for
the run with acid to alcohol molar ratio of 1:4 is approximately
17% more than that for the run with a ratio of 1:2 and 32%
more than that for the run with a ratio of 1:1.

Dakshinamurty

et al. (1984)

, who studied the esterification reaction of 1-propa-

nol with propionic acid over Dowex-50W, found that the con-
version of acid increased with increasing initial alcohol to acid
molar ratio. Also, some other workers (

Yadav and Kulkarni,

2000; Yadav and Thathagar, 2002

) found that the conversion

of acid increases with decreasing the acid to alcohol molar ra-
tio.

Fig. 10

is a re-plot of

Fig. 9

but in terms of conversion

of the limiting reactant rather than propionic acid. As shown
in

Fig. 10

, changing the molar ratio from 1:1 to 1:2 increases

the conversion of the limiting reactant. A further increase in
the amount of alcohol initially leads to a higher conversion
of the limiting reactant. Also, increasing the molar ratio from
1:1 to 2:1 increases the conversion of the limiting reactant
and a further increase in the acid to alcohol initial molar ra-
tio leads to a more significant increase in the conversion of the

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3209

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

1 : 1

1 : 2

2 : 1

4 : 1

1 : 4

Fig. 9. Effect of propionic acid to 1-propanol molar ratio on the conversion
of propionic acid at 323 K, 900 rpm and 40 g dry cat/L catalyst loading of
Dowex 50Wx8-400.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of the limiting reactant

1 : 1

1 : 2

1 : 4
2 : 1
4 : 1

Fig. 10. Effect of propionic acid to 1-propanol molar ratio on the conversion
of the limiting reactant at 323 K, 900 rpm and 40 g dry cat/L catalyst loading
of Dowex 50Wx8-400.

limiting reactant.

Fig. 10

shows that acid to alcohol molar ratio

of 2:1 gives a higher conversion of the limiting reactant than
1:2. Also, a more significant increase in conversion was ob-
tained at a ratio of 4:1 rather than 1:4. As a result, conversion
of the limiting reactant is higher in cases different than 1:1 and
a more significant increase in conversion of the limiting reac-

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

run with water (acid : alcohol : water = 1 : 1 : 1)
run without water (acid : alcohol = 1 : 1)

Fig. 11. Effect of adding water initially on the conversion of propionic
acid at 323 K, 900 rpm, 1:1 propionic acid to 1-propanol molar ratio and
40 g dry cat/L catalyst loading of Dowex 50Wx8-400.

tant is obtained by increasing the initial amount of acid than
increasing the initial amount of alcohol.

4.3.7. Effect of adding water to the reactants mixture

The effect of adding water to the reactant mixture on the

conversion of propionic acid is shown in

Fig. 11

. A compari-

son between two experiments was made over a time period of
4 h, under the same conditions. However, one experiment was
done in the presence of initial amounts of water in the reaction
mixture (acid: alcohol: water equals 1:1:1). It is obvious from

Fig. 11

that adding water decreased the conversion of propionic

acid. This observed decrease is due to several reasons. Water
is one of the products for this esterification reaction, so adding
water initially will promote the reverse reaction. In addition,
water is found to have a very strong affinity for cation-exchange
resins such as Dowex 50Wx8-400 (

Pöpken et al., 2000

). Also,

adsorption of water is most likely higher than the other species
present in the system.

Song et al. (1998)

studied methyl acetate

esterification over Amberlyst-15 and measured the adsorption
of the different species present in their system relative to the
adsorption of methyl acetate. They found that water is adsorbed
more strongly than the other reactants and products, which fur-
ther support our results. Therefore, the initial presence of water
in the system will lead to lower conversions.

4.3.8. Effect of using different acids

The study of the esterification of 1-propanol with different

acids (acetic, butyric and propionic) is represented in

Fig. 12

.

This figure shows that the conversion of propionic acid de-
creased with increasing the acid chain length due to steric hin-
drance. It was found that at 4 h of reaction, the conversion of
acetic acid was 62%, propionic acid 48% and butyric acid 31%.

background image

3210

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of acid

Propionic acid

Acetic acid

Butyric acid

Fig. 12. Effect of using different acids on the conversion of the acid at 323 K,
900 rpm, 1:1 acid to 1-propanol molar ratio and 40 g dry cat/L catalyst loading
of Dowex 50Wx8-400.

This implies that the conversion and hence the reaction rate in-
creases as the chain length decreases. A similar conclusion was
made by

Lilja et al. (2002)

.

4.3.9. Effect of using different alcohols

The effect of different alcohols on the conversion of pro-

pionic acid was investigated by running multiple experiments
using different alcohols (methanol, ethanol, 1-propanol and 1-
butanol) and holding other parameters constant. Increasing the
chain length of the alcohol decreased the conversion of propi-
onic acid, as indicated in

Fig. 13

. The propionic acid conversion

at 4 h of reaction was 67%, 55%, 48%, 46% when reacting with
methanol, ethanol, 1-propanol and 1-butanol, respectively. The
hindrance effect is the main reason for the decrease in the con-
version of propionic acid as the alcohol chain length increases.

Awad et al. (1997)

and

Lilja et al. (2002)

results agree with our

results. These authors concluded that alcohol chain length had
a retarding effect on the overall reaction rate. The esterification
of propionic acid with 1-propanol and 2-propanol over Dowex
50Wx8-400 was investigated and the results are shown in

Fig. 14

. As shown in

Fig. 14

, the reaction of propionic acid with

1-propanol had a higher conversion than the reaction of the
propionic acid with 2-propanol. At 4 h of reaction, it was found
that the conversion of propionic acid reacting with 1-propanol
is 48% while the conversion of propionic acid reacting with
2-propanol is 10%. This indicates that branching of the alcohol
had a retarding effect on the conversion and hence the reaction
rate due to steric hindrance.

Awad et al. (1997)

found that

the rate of esterification reaction of propionic acid with linear
alcohols was higher than that of branched ones.

Lilja et al.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

methanol

ethanol

1-propanol

1-butanol

Fig. 13. Effect of using different alcohols on the conversion of propionic acid
at 323 K, 900 rpm, 1:1 propionic acid to alcohol molar ratio and 40 g dry cat/L
catalyst loading of Dowex 50Wx8-400.

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

0

3600

7200

10800

14400

18000

Time (sec)

Conversion of propionic acid

2-propanol

1-propanol

Fig. 14. Effect of using 2-propanol versus 1-propanol on the conversion of
propionic acid at 323 K, 900 rpm, 1:1 propionic acid to alcohol ratio and
40 g dry cat/L catalyst loading of Dowex 50Wx8-400.

(2002)

, who studied the esterification of acetic and propionic

acids with different alcohols over Smopex-101, as a catalyst,
demonstrated that branching of the alcohol chain decreased
the reaction rate. In contrast,

El-Noamany et al. (1994)

found

that branching in the hydrocarbon chain length of an aliphatic
primary alcohol had insignificant effect on the esterification

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3211

Table 6
The UNIFAC groups present in the different components and their R and Q
values

Group

CH

3

CH

2

COOH

OH

H

2

O

CH

2

COO

Propionic acid

1

1

1

0

0

0

1-Propanol

1

2

0

1

0

0

Propyl propionate

2

2

0

0

0

1

Water

0

0

0

0

1

0

R

0.9011

0.6744

1.3013

1

0.92

1.6764

Q

0.848

0.54

1.224

1.2

1.4

1.42

reaction of acetic acid with monohydric alcohols (1-butanol,
2-butanol, 1-hexanol, cyclo-hexanol and benzyl alcohol) using
Amberlite IR-120 as a catalyst.

4.4. Modeling

From the different kinetic runs involved in this investiga-

tion, 10 runs were selected for modeling. These runs (4, 10,
11 and 15–21) represented the reaction of propionic acid with
1-propanol catalyzed by Dowex 50Wx8-400 at different re-
action conditions. Kinetic data from these selected runs are
fitted to 40 rate expressions as discussed previously. The activ-
ity coefficients of the reactants and the products used in these
equations were determined using the UNIFAC model. In this
investigation, UNIFAC was incorporated into the used Mathe-
matica software. To calculate the UNIFAC activity coefficients
in our system, the four components present in the system were
divided into the following sub-groups as reported in

Table 6

:

propionic acid (1 CH

3

, 1 CH

2

and 1 COOH); 1-propanol (1

CH

3

, 2 CH

2

and 1 OH); 1-propyl propionate (2 CH

3

, 2 CH

2

and 1 CH

2

COO); water (1 H

2

O).

Experimental data fitted to 40 rate expressions were tested

using Mathematica software (Statistics Nonlinear fit function).
The aim of this fitting is to minimize the mean-square differ-
ences between calculated values of rate (r

calc

obtained from

Mathematica) with values obtained from experimental data
(r

exp

) as shown below:

min

=

all data samples

(r

calc

r

exp

)

2

,

(22)

where the experimental data of mole fraction of limiting com-
ponent were fitted as a function of time (x

exp

=f (t)). Then this

function was differentiated in order to obtain the experimental
rate (r

exp

) as follows:

r

exp

= −n

dx

exp

dt

.

(23)

This data-fitting was done to obtain the model parameters. Some
of these models gave negative values for the adsorption equi-
librium constants (K

acid

, K

alc

, K

ester

and/or K

water

) and/or the

pre-exponential factor (A

f

). As a result, out of 40 models only

17 models might be suitable for describing the reaction kinet-
ics.

Tables 7–9

list the number and five parts name assigned

for these 40 models. The first part is for the number of sites
involved, either dual (Dl) or single (Sg). The second part is for

T

able

7

Kinetic

parameters

for

the

P-H

and

the

dif

ferent

dual-site

models

Model

#

Model

A

f

E

f

k

a

(mol/g/s)

K

acid

K

alc

K

ester

K

w

ater

T.

Av

g

.

(mol/g/s)

(J/mol)

Error

303.15

K

313.15

K

323.15

K

333.15

K

1P

H

2

.61E

+

06

66,342

9.7E

06

2.2E

05

4.9E

05

1.0E

04

2.26

2

Dl-Ac

Alc-Ads-Ac-

13

.16E

+

08

83,557

1.3E

06

3.6E

06

9.8E

06

2.5E

05

2.00

0.99

0.00

0.41

3

Dl-Ac

Alc-Ads-Ac-

28

.60

E

+

08

86,279

1.2E

06

3.5E

06

9.7E

06

2.5E

05

4.00

0.98

0.01

1.09

4

Dl-Ac

Alc-Ads-Ac-

31

.09E

+

09

86,955

1.1E

06

3.4E

06

9.6E

06

2.5E

05

3.00

0.99

0.10

1.91

5

Dl-Ac

Alc-Ads–Alc-

14

.93E

+

09

91,422

8.7E

07

2.8E

06

8.2E

06

2.3E

05

1.05

5.30

14.93

10

6

Dl-Ac

Alc-Ads-Alc-

28

.71E

+

20

159,373

3.0E

07

2.3E

06

1.5E

05

8.9E

05

2

6

.59E

+

22

1.00

10.04

7

Dl-Ac

Alc-Ads–Alc-

38

.50

E

+

10

99,342

6.5E

07

2.3E

06

7.4E

06

2.3E

05

1.09

1.67

0.99

9.80

8

Dl-Ac

Alc-Rxn-Ac

Alc-

17

.12E

+

06

65,720

3.4E

05

7.8E

05

1.7E

04

3.5E

04

1.00

1.57

0.90

2.00

4.33

9

Dl-Ac

Alc-Rxn-Ac

Alc-

21

.05E

+

07

68,613

1.6E

05

3.8E

05

8.5E

05

1.8E

04

2.60

3.00

1.50

6.00

2.63

10

Dl-Ac

Alc-Rxn-Ac

Alc-

31

.15E

+

07

68

549

1.8E

05

4.2E

05

9.5E

05

2.0E

04

1.80

2.48

0.70

6.28

1.83

11

Dl-Ac

Alc-Des-Wt-

13

.42E

+

05

74,676

4.6E

08

1.2E

07

2.9E

07

6.7E

07

0.13

0.37

0.76

0.03

12

Dl-Ac

Alc-Des-Wt-

25

.57E

+

04

57,390

7.2E

06

1.5E

05

2.9E

05

5.6E

05

2.00

3.00

1.30

6.00

5.78

13

Dl-Ac

Alc-Des-Wt-

33

.01E

+

05

70,064

2.5E

07

6.2E

07

1.4E

06

3.1E

06

0.25

0.25

1.372

0.13

14

Dl-Ac

Alc-Des-Es-

13

.42E

+

05

70,000

3.0E

07

7.2E

07

1.7E

06

3.6E

06

0.21

0.46

0.17

1.08

15

Dl-Ac

Alc-Des-Es-

2

v

e

82,915

––––

0.04

0.88

5.58E

12

0.38

16

Dl-Ac

Alc-Des-Es-

3

2.32E

01

50,606

4.4E

10

8.4E

10

1.5E

09

2.7E

09

0.37

0.13

0.00

1.30

a

Reaction,

adsorption

or

desorption.

background image

3212

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

T

able

8

Kinematic

parameters

for

the

dif

ferent

single-site

models

Model

Model

A

f

E

f

k

a

(mol/g/sec)

K

acid

K

alc

K

ester

K

w

ater

T.

Av

g

.

#

(mol/g/s)

(J/mol)

Error

303.15

K

313.15

K

323.15

K

333.15

K

17

Sg-Ac-Ads-Ac-

17

.33E

+

11

103,119

1.2E

06

4.6E

06

1.6E

05

5.0E

05

8

.04E

+

13

1.94

18

Sg-Ac-Ads-Ac-

29

.44E

+

12

110,217

9.6E

07

3.9E

06

1.4E

05

4.9E

05

2.33

3.82

11.37

19

Sg-Ac-Ads-Ac-

31

.42E

+

13

111,388

9.1E

07

3.7E

06

1.4E

05

4.9E

05

2

14.04

20

Sg-Ac-Rxn-Ac-

14

.22E

+

07

75,089

4.9E

06

1.3E

05

3.1E

05

7.1E

05

5.11

6.00

6.33

21

Sg-Ac-Rxn-Ac-

2

v

e

66,223

––––

0.23

0.57

22

Sg-Ac-Rxn-Ac-

3

v

e

65,441

––––

0.21

0.64

23

Sg-Ac-Des-Wt-

12

.72E

+

07

82,208

1.9E

07

5.3E

07

1.4E

06

3.5E

06

0.11

0.16

4.63

24

Sg-Ac-Des-Wt-

27

.94E

+

06

69,154

9.64E

06

2.31E

05

5.27E

05

1.14E

04

3.09

10.00

6.66

25

Sg-Ac-Des-Wt-

3

v

e

42,844

––––

1

.87E

+

10

7

.31E

+

10

26

Sg-Ac-Des–Es–

17

.33E

+

05

72,264

2.59E

07

6.47E

07

1.53E

06

3.42E

06

0.3

0.22

5.64

27

Sg-Ac-Des-Es-

22

.09E

+

06

82,760

1.15E

08

3.27E

08

8.74E

08

2.20E

07

0.076

0.013

5.71

28

Sg-Ac-Des-Es-

3

v

e

82,761

––––

0.03

0.15

29

Sg-Alc-Ads-Alc-

11

.85E

+

14

118,640

6.7E

07

3.0E

06

1.2E

05

4.6E

05

0.62

1.25

12.61

30

Sg-Alc-Ads-Alc-

21

.53E

+

15

124,541

5.3E

07

2.6E

06

1.1E

05

4.5E

05

2.00

3.59

13.32

31

Sg-Alc-Ads-Alc-

33

.29E

+

15

126,723

4.8E

07

2.4E

06

1.1E

05

4.4E

05

7

.00

E

+

17

9.17

32

Sg-Alc-Rxn-Alc-

17

.77E

+

06

66,939

2.3E

05

5.3E

05

1.2E

04

2.5E

04

0.55

0.70

2.05

33

Sg-Alc-Rxn-Alc-

28

.00

E

+

06

67,322

2.0E

05

4.7E

05

1.0E

04

2.2E

04

0.60

0.80

1.65

34

Sg-Alc-Rxn-Alc-

35

.95E

+

06

67,607

1.3E

05

3.1E

05

7.0E

05

1.5E

04

1.00

1.35

1.77

35

Sg-Alc-Des-Wt-

1

v

e

59,723

––––

1

.52E

+

08

3

.82E

+

08

36

Sg-Alc-Des-Wt-

2

v

e

55,561

––––

2

.32E

+

08

7

.24E

+

08

37

Sg-Alc-Des-Wt-

3

v

e

54,002

––––

6

.22E

+

08

2

.44E

+

09

38

Sg-Alc-Des-Es-

1

v

e

57,298

––––

2

.24E

+

08

5

.65E

+

08

39

Sg-Alc-Des-Es-

22

.20

E

+

06

82,715

1.2E

08

3.5E

08

9.4E

08

2.4E

07

0.08

0.01

4.71

40

Sg-Alc-Des-Es-

3

v

e

82,757

––––

0.03

0.15

a

Reaction,

adsorption

or

desorption.

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3213

Table 9
Errors obtained while fitting experimental runs with different models

Model #

Run no.

4

10

11

15

16

17

18

19

20

21

T.Avg.Error

Catalyst loading (g dry cat/L)

40

20

60

40

40

40

40

40

40

40

Temperature (K)

323.15

323.15

323.15

303.15

313.15

333.15

323.15

323.15

323.15

323.15

Acid:alcohol rounded molar ratio

1:1

1:1

1:1

1:1

1:1

1:1

2:1

1:2

1:4

4:1

Models tested

Individual error of each experimental run

1

PH

1.07

1.74

2.17

1.21

1.06

2.39

1.05

5.54

5.10

1.28

2.26

8

Dl-Ac Alc-Rxn-Ac Alc-

1

6.19

3.62

10.62

0.98

1.15

9.52

0.91

3.06

6.01

1.27

4.33

9

Dl-Ac Alc-Rxn-Ac Alc-

2

3.44

1.76

7.30

1.12

0.99

5.54

0.81

2.17

2.32

0.86

2.63

10

Dl-Ac Alc-Rxn-Ac Alc-

3

1.13

1.03

4.14

1.39

1.35

1.93

2.43

3.37

1.18

0.37

1.83

12

Dl-Ac Alc-Des-Wt-

2

4.61

6.16

4.37

2.38

2.80

5.45

1.70

5.99

22.67

1.64

5.78

18

Sg-Ac-Ads-Ac-

2

12.35

8.78

17.62

7.49

9.32

3.72

1.72

17.75

29.79

5.14

11.37

20

Sg-Ac-Rxn-Ac-

1

6.96

4.71

11.09

3.17

2.36

6.96

4.32

4.77

14.22

4.77

6.33

23

Sg-Ac-Des-Wt-

1

4.23

3.47

5.80

3.49

2.63

0.74

1.03

6.05

18.09

0.78

4.63

24

Sg-Ac-Des-Wt-

2

4.42

6.45

2.57

3.86

4.16

3.11

1.98

8.87

29.43

1.77

6.66

26

Sg-Ac-Des-Es-

1

3.59

4.35

3.04

3.14

2.50

4.66

1.30

8.45

24.26

1.12

5.64

27

Sg-Ac-Des-Es-

2

5.02

4.09

7.14

4.14

3.54

0.43

0.72

6.64

24.17

1.22

5.71

29

Sg-Alc-Ads-Alc-

1

14.37

10.54

18.97

8.06

11.00

1.49

10.30

6.96

36.12

8.26

12.61

30

Sg-Alc-Ads-Alc-

2

14.65

10.71

19.78

8.33

11.64

0.72

10.41

6.59

42.01

8.38

13.32

32

Sg-Alc-Rxn-Alc-

1

2.68

2.20

5.00

1.22

0.75

2.53

1.26

1.81

2.36

0.67

2.05

33

Sg-Alc-Rxn-Alc-

2

1.49

1.72

2.99

1.35

0.93

1.18

2.08

3.00

1.36

0.40

1.65

34

Sg-Alc-Rxn-Alc-

3

0.91

1.74

2.27

1.54

0.80

2.17

3.00

2.15

2.63

0.53

1.77

39

Sg-Alc-Des-Es-

2

4.18

3.40

5.94

3.44

2.94

0.35

0.67

5.37

19.68

1.09

4.71

the reactants being adsorbed; acid (Ac) and/or alcohol (Alc).
The third part is for the controlling step; adsorption (Ads), sur-
face reaction (Rxn) or desorption (Des). The fourth part is for
the controlling component. The fifth part is for the correction
factor for the resin affinity for water (

= 1, 2 or 3). The total

average error (T.Avg.Error) between the predicted and experi-
mental mole fractions of acid was calculated as follows:

T.Avg.Error

=

all data samples

|x

exp

x

pred

|/x

exp

n

samples

× 100.

(24)

Tables 7–9

show the total average error obtained for these 17

models which ranged between 1.65% and 13.32%. For all of
these models (except the P-H model), the same trend of adsorp-
tion equilibrium constant exists K

water

> K

alc

> K

acid

> K

ester

(see

Tables 7 and 8

). The trend of our calculated adsorption

equilibrium constants agrees with the trend in solubility pa-
rameter of the components as has been suggested in our earlier
work (

Ali and Merchant, 2006

). The solubility parameter val-

ues obtained from the AICHE DIPPER

at 25

C are 47.81,

24.45, 19.49 and 17.57 (J/cm

3

)

0.5

for water, 1-propanol, propi-

onic acid and 1-propyl propionate, respectively.

Table 9

shows

that there are only four models (models # 10, 32, 33 and 34)
which perform better than the simplest possible model namely
the P-H model (model # 1 with T. Avg. Error of 2.26%). These
four models have total average error values below 2.26%. All
these models are surface reaction controlling models with dif-
ferent reaction mechanisms. The mechanism of three of these
models is a single site (E–R and M-E–R) where 1-propanol is
being adsorbed reacting with non-adsorbed propionic acid re-
sulting in adsorbed water and non-adsorbed ester with different
water adsorption affinities (

= 1, 2 or 3), and the mechanism

for the other model is a dual site (M-L–H) where both propi-
onic acid and 1-propanol are being adsorbed with a water ad-
sorption affinity of 3 (

= 3). The total average errors of these

four models are 2.05%, 1.65%, 1.77% and 1.83%, respectively.
The activation energy (E

f

) for the five potential models varied

between 66.3 and 68.5 kJ/mol while the pre-exponential factor
(A

f

) varied between 2.61E + 06 and 1.15E + 07 mol/g/s.

To select the best model representing the data, the individual

errors (as reported in

Table 9

) for each model were compared

in addition to the accuracy of prediction of equilibrium mole
fractions by each model. The equilibrium mole fractions were
obtained from the kinetic models by using the fitted parameters
to predict the mole fractions at large enough times where the
system is considered to be in an equilibrium status. As shown
in

Table 9

, the P-H model had an error around 5% in two

runs (runs # 19 and 20). The M-L–H (model # 10), where the
surface reaction is rate controlling and the water adsorption
affinity is 3 (

= 3), gave a bad prediction of the equilibrium

mole fractions. The three remaining models (models # 32, 33
and 34) are single-site reaction mechanism models where 1-
propanol is being adsorbed with the surface reaction being the
rate controlling step. The difference between these models is
the correction term representing the water adsorption affinity
(

= 1, 2 or 3). All three models give a good prediction for

the equilibrium mole fraction at different reaction temperatures
with errors less than 1.5%.

Since one of the aims of this work is to arrive at the reac-

tion mechanism which best describes the kinetics of this het-
erogeneous system over the wide range of conditions studied
(this includes different temperatures, acid to alcohol molar ra-
tios and catalyst loadings), the standard deviation of the indi-
vidual error of each experimental run around the total average

background image

3214

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

0

0.1

0.2

0.3

0.4

0.5

0

3600

7200

10800

14400

18000

Time (sec)

Mole fraction of propionic acid

Exp.; 20 g dry cat/L

Sg-Al-Rxn-Al-a2; 20g dry cat/L

Exp.; 40 g dry cat/L

Sg-Al-Rxn-Al-a2; 40g dry cat/L

Exp.; 60 g dry cat/L

Sg-Al-Rxn-Al-a2; 60g dry cat/L

Fig. 15. Experimental versus predicted (by the best model; model # 33) mole
fraction of propionic acid as a function of time at different catalyst loadings
of Dowex 50Wx8-400 at 323 K, 900 rpm, 1:1 propionic acid to 1-propanol
molar ratio.

error (STDEV) were performed for models # 32, 33 and 34
according to the following equation:

STDEV

=

(Individ.Error − T.Avg.Error)

2

(N − 1)

,

(25)

where Individ.Error is the individual error of each experimental
run, T.Avg.Error is the total average error, and N is the number
of experimental runs considered in the modeling process.

STDEV for models # 32, 33 and 34 were found to be equal

to 1.3, 0.8 and 0.8, respectively. This indicates that models #
33 and 34 are better than model # 32 in predicting reaction
kinetics. Though the standard deviation values of models # 33
and 34 are comparable to one another, the total average error
of model # 33 is lower than that of model # 34, 1.65% versus
1.77%, respectively. Hence model # 33, a single-site M-E–R
model wherein adsorbed 1-propanol reacts with non-adsorbed
propionic acid resulting in non-adsorbed ester and adsorbed
water (with

= 2), is selected as the model best capable of

describing the studied reaction kinetics.

Interestingly, earlier workers (

Dakshinamurty et al., 1984;

Liu and Tan, 2001

) have also proposed single-site mechanisms

for acidic ion-exchange resin catalyzed esterifications involv-
ing propionic acid.

Dakshinamurty et al. (1984)

, who studied

the propyl propionate esterification reaction over Dowex-50W,
proposed that the rate controlling step was the surface reac-
tion (involving a single site) between adsorbed propionic acid
reacting with 1-propanol in the bulk. However, no such mathe-
matical expression was presented; but rather, an empirical rela-
tionship correlating the specific reaction rate constant in terms
of the studied variables was reported. Furthermore,

Liu and Tan

(2001)

found that their heterogeneously catalyzed esterification

reaction over Amberlyst 35 follows the E–R theory in which

0

0.1

0.2

0.3

0.4

0.5

0

3600

7200

10800

14400

18000

Time (sec)

Mole fraction of propionic acid

Exp.; 303.15 K

Sg-Al-Rxn-Al-a2; 303.15K

Exp.; 313.15 K

Sg-Al-Rxn-Al-a2; 313.15K

Exp.; 323.15 K

Sg-Al-Rxn-Al-a2; 323.15K

Exp.; 333.15 K

Sg-Al-Rxn-Al-a2; 333.15K

Fig. 16. Experimental versus predicted (by the best model; model # 33) mole
fraction of propionic acid as a function of time at different temperatures, cat-
alyst loading 40 g dry cat/L of Dowex 50Wx8-400, 900 rpm and 1:1 propionic
acid to 1-propanol molar ratio.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0

3600

7200

10800

14400

18000

Time (sec)

Mole fraction of propionic acid

Exp. Ratio; 2 : 1

Sg-Al-Rxn-Al-a2; 2 : 1

Exp. Ratio; 1 : 2

Sg-Al-Rxn-Al-a2; 1 : 2

Exp Ratio; 1 : 1

Sg-Al-Rxn-Al-a2; 1 : 1

Exp. Ratio; 1 : 4

Sg-Al-Rxn-Al-a2; 1 : 4

Exp. Ratio; 4 :1

Sg-Al-Rxn-Al-a2; 4 : 1

Fig. 17. Experimental versus predicted (by the best model; model # 33) mole
fraction of propionic acid as a function of time at different propionic acid to
1-propanol molar ratios at 323 K, catalyst loading of 40 g dry cat/L of Dowex
50Wx8-400 and 900 rpm.

adsorbed propionic acid reacts with 1-butanol in the bulk. They
(

Liu and Tan, 2001

) assumed that the rate controlling step is the

adsorption step involving the acid species. However, there are
some objections regarding the values of the reported adsorption
equilibrium constants for propionic acid and water, which were
fitted as a function of temperature. One of these objections re-
garding the zero values of the adsorption equilibrium constants
for propionic acid and water at 353 K, which are not reasonable.
Another one is related to the un-explained and non-systematic
variations and changes in the values of the propionic acid

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3215

adsorption equilibrium constants as a function of temperature.
In addition, for the temperature range from 363 to 383 K, they
(

Liu and Tan, 2001

) reported that the adsorption equilibrium

constants for propionic acid are higher than that for water,
which is opposite to what is expected.

Song et al. (1998)

ex-

perimentally determined the adsorption equilibrium constants
for the species involved in methyl acetate synthesis using Am-
berlyst 15. They (

Song et al., 1998

) found that the adsorp-

tion equilibrium constant for water was higher than acetic acid.
Also,

Pöpken et al. (2000)

found the same to be true for the

same heterogeneously catalyzed system. In addition, in our pre-
vious work (

Ali and Merchant, 2006

) we have found that the

adsorption equilibrium constants for the different species are
proportional to their solubility parameters (as found in this cur-
rent study and mentioned previously), which implies that the
adsorption constant for water is expected to be higher in value
than for propionic acid. Furthermore, both

Liu and Tan (2001)

and

Dakshinamurty et al. (1984)

suggest mechanisms wherein

the acid is selectively adsorbed on the catalyst and the alcohol
remains in the bulk phase. This scenario is not plausible con-
sidering the fact that the catalysts involved are sulfonic acid
ion-exchange resins and alcohols are known to be apprecia-
bly sorbed on such catalysts (

Rehfinger and Hoffmann, 1990;

Song et al., 1998; Pöpken et al., 2000

). Moreover, the model

proposed by

Dakshinamurty et al. (1984)

which is represented

by model # 20 (Sg-Ac-Rxn-Ac-

1; single-site acid adsorbed-

surface reaction controlled, with

equal to 1) in this investiga-

tion showed a total average error of 6.33%. This error is signifi-
cantly higher than for model # 33 for propyl propionate synthe-
sis over Dowex 50Wx8-400 suggesting that the latter model is
better in predicting the reaction kinetics at different conditions.

The activation energy for the forward reaction was estimated

to be 67.3 kJ/mol. This value seems reasonable since it is quite
comparable (even though it is for a different catalyst) to the
value of 63.7 kJ/mol reported by

Lee et al. (2002)

for the es-

terification of propionic acid and 1-butanol over Amberlyst 35,
which was best represented by L–H model. On the other hand,
the value of the activation energy of 39.1 kJ/mol reported by

Dakshinamurty et al. (1984)

for the esterification reaction of 1-

propanol and propionic acid over Dowex-50W is significantly
lower than our value. In fact, this value (39.1 kJ/mol) is sig-
nificantly smaller than the reported value (activation energy of
54.3 kJ/mol) for even less hindered esterification system stud-
ied by

Krishnaiah and Rao (1984)

for propyl acetate synthesis

by Dowex-50W.

The best model results were compared with the experimental

values for different reaction conditions in

Figs. 15–17

. These

figures show comparisons between the experimental and the
predicted (by the best model; model # 33) mole fraction of
propionic acid at different catalyst loading, temperature and
acid to alcohol molar ratio. The M-E–R model (model # 33) was
effective in predicting the mole fraction of propionic acid for the
reaction system under different reaction conditions as shown
in

Figs. 15–17

. In addition, the mole fraction of propionic acid

was accurately predicted by model # 33 as can be seen in

Fig. 18

. This parity plot clearly goes along with our finding

(low total average error of 1.65% with STDEV of 0.8).

0

0.2

0.4

0.6

0.8

1

0

0.2

0.4

0.6

0.8

1

Calculated propionic acid mole fraction

Experimental propionic acid mole fraction

Fig. 18. Parity plot for the experimental versus calculated (using the best
model; model # 33) propionic acid mole fraction.

An attempt was made to further improve the prediction ca-

pability of the models by considering the correction factor

as

a variable rather than a constant value. However, it was found
that the model prediction by M-E–R model gave the best pre-
diction at

= 2 and gave slightly less accurate predictions for

other values of

ranging between 1 and 3. Therefore, it seems

that the optimum value of

is 2 as presented by model # 33.

5. Conclusions

The esterification reaction of propionic acid with 1-propanol

was successfully carried out over Dowex 50Wx8-400. The
experimental results indicated that, under the implemented
reaction conditions, Dowex 50Wx8-400 was a more effective
catalyst for this reaction compared with the other ion-exchange
resins tested, namely Amberlite IR-120 and Amberlyst 15.
A stirrer speed of 200 rpm or higher, in the used 1 L reactor,
was found to be effective for eliminating external diffusion ef-
fect. The significance of internal mass transfer limitation was
quantified by the use of the Weisz–Prater criterion for exper-
iments carried at different catalyst concentration, temperature
and acid to alcohol molar ratio. It was found that the internal
diffusion effect was negligible (C

W P

>1) under the employed

esterification reaction conditions. The conversion of acid in-
creased with increasing temperature and catalyst loading and
decreased with increasing initial amount of acid. The increase
in the acid chain length had a retarding effect on the conver-
sion of acid. Similarly, an increase in alcohol chain length or
branching decreased conversion. Adding water initially to acid
and alcohol mixture also decreased the conversion. The stan-
dard enthalpy change of reaction was found to be 6.4 kJ/mol.
This esterification reaction was found to be mildly exothermic.
Among various kinetic models employed in this investigation,

background image

3216

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

the reaction of propionic acid with 1-propanol catalyzed by
Dowex 50Wx8-400 was best presented by the modified M-E–R
model with a total average error of 1.65%. This model is based
on a controlling step of surface reaction between adsorbed 1-
propanol with non-adsorbed propionic acid forming 1-propyl
propionate and water (with a correction factor for the resin
affinity for water equal to 2). The activation energy for the for-
ward reaction was estimated to be 67.3 kJ/mol. UNIFAC was
used successfully to account for the non-ideal thermodynamic
behavior of the reactants and the products.

Notation

a

i

activity of component i in the liquid phase

a

mn

a parameter of the interaction coefficient be-
tween groups m and n

A

f

pre-exponential factor for the forward reac-
tion leading to ester formation, mol/g/s

b

mn

a parameter of the interaction coefficient be-
tween groups m and n

C

li

limiting reactant concentration in the mixture
at a given time, mol/cm

3

C

W P

Weisz–Prater parameter

D

e

effective diffusivity, cm

2

/s

D

li

diffusivity of limiting reactant in component
i, cm

2

/s

D

lm

diffusivity of limiting reactant in the mixture,
cm

2

/s

E–R

Eley–Rideal model

E

f

activation energy for the forward reaction
leading to ester formation, kJ/mol

H

0

R

standard enthalpy change of reaction, kJ/mol

Individ.Error

individual error of each experimental run

k

f

forward reaction rate constant for esterifica-
tion, mol/g/s

k

forward reaction rate constant (adsorption,
desorption or surface reaction), mol/g/s

K

a

esterification reaction equilibrium constant
for the overall reaction

K

i

adsorption equilibrium constant for species i
present in the system

K

s

experimentally measured esterification reac-
tion equilibrium constant

K

SR

surface reaction equilibrium constant at a ref-
erence temperature

L–H

Langmuir–Hinshelwood model

M-E–R

modified Eley–Rideal model

M-L–H

modified Langmuir–Hinshelwood model

M

cat

mass of the catalyst, g

M

i

molar mass of component i, g/mol

n

total number of moles in the system, mol

n

samples

number of samples

N

total number of runs considered for modeling

P-H

pseudo-homogeneous model

Q

UNIFAC group area parameter

r

catalyst pellet radius, cm

r

A(obs)

observed reaction rate at a given time, mol/g
of catalyst/s

r

calc

calculated reaction rate, mol/s

r

exp

reaction rate determined from experimental
data, mol/s

R

UNIFAC group volume parameter

R

c

ratio of catalyst pellet volume to catalyst pel-
let external surface area, cm

R

2

fit

correlation coefficient for a fit

R

g

ideal gas law constant, 8.314 J/mol/K

STDEV

standard deviation

t

time, s

T

temperature, K

T.Avg.Error

total average error

T

R

reference temperature, K

V

i

molar volume of component i, cm

3

/mol

x

exp

experimental mole fraction

(x

i

)

eq

equilibrium mole fraction of component i

x

pred

predicted mole fraction

z

coordination number

Greek letters

exponential term accounting for water affin-
ity for the resin

i

activity coefficient of component i

C

i

combinatorial part of the activity coefficient
of component i

R

i

residual part of the activity coefficient of
component i

i

viscosity of component i, cp

m

viscosity of the mixture, cp

surface area fraction

v

void fraction of the catalyst

c

catalyst density, g/cm

3

mn

energy interaction parameter between com-
ponents m and n

k

activity coefficient of group k at mixture com-
position

i

k

activity coefficient of group k of pure com-
ponent i

volume fraction

References

Ali, S.H., Merchant, S.Q., 2006. Kinetics of the esterification of acetic

acid with 2-propanol: impact of different acidic cation exchange resins
on reaction mechanism. International Journal of Chemical Kinetics 38,
593–612.

Awad, M.M., Salem, A.M., Swelam, A.A., 1997. Kinetics of catalyzed

esterification of propionic acid with various alcohols using synthetic cation
exchange resin. Journal of the Indian Chemical Society 74, 467–469.

Bart, H.J., Kaltenbrunner, W., Landschutzer, H., 1996. Kinetics of

esterification of acetic acid with propyl alcohol by heterogeneous catalysis.
International Journal of Chemical Kinetics 28, 649–656.

Bhatia, S., Rajamani, K., Rajkhowa, P., Rao, M.G., 1973. Ion exchange resin

catalysis. Ion Exchange and Membrane 1, 127–133.

Chang, J.R., Yeh, M.Y., 1984. Kinetics of phase transfer catalytic preparation

of benzyl benzoate. Journal of the Chinese Chemical Society 31, 185–190.

background image

S.H. Ali et al. / Chemical Engineering Science 62 (2007) 3197 – 3217

3217

Chiplunkar, M., Hong, M., Malone, M.F., Doherty, M.F., 2005. Experimental

study of feasibility in kinetically-controlled reactive distillation. A.I.Ch.E.
Journal 51 (2), 464–479.

Dakshinamurty, P., Ramarao, M.V.S., Ramachandramurty, Ch.V., 1984.

Kinetics of catalytic esterification of propan-1-ol with propanoic acid using
cation exchange resin. Journal of Chemical Technology and Biotechnology
34A, 257–261.

El Ewady, Y.A., El Wakil, A.M., El Nahas, M.R., Moussa, M.N.H., 1984.

Esterification of methanol with a homologous series of acids using
Amberlite IR-20. Journal of the Indian Chemical Society LXI, 517–519.

El-Noamany, H.M., Shaheen, F.A., El-Kenawy, O.S., Zaher, F.A., 1994.

Studies on some factors affecting the rate of esterification of acetic acid
with monohydric alcohols. Modeling Measurements and Control 43 (1),
25–43.

Fogler, H.S., 1992. Elements of Chemical Reaction Engineering. second ed..

Prentice-Hall, New York. pp. 607–647.

Gangadwala, J., Manker, S., Mahajani, S., 2003. Esterification of acetic acid

with butanol in presence of ion exchange resins as catalysts. Industrial
and Engineering Chemistry Research 42, 2146–2155.

Hansen, H.K., Coto, B., Kuhlmann, B., 1992. UNIFAC with linearly

temperature-dependant group-interactional parameters. Technical Report
(SEP 9212), IVC-SEP Research Engineering Center, Institute for
Kemiteknik, The Technical University of Denmark, Lyngby.

Krishnaiah, D., Rao, M.B., 1984. Kinetics of esterification of n-propyl alcohol

with acetic acid catalyzed by Dowex-50W. Indian Journal of Technology
22, 268–271.

Lee, M.J., Wu, H.T., Kang, C.H., Lin, H.M., 1999. Kinetic behavior of amyl

acetate synthesis catalyzed by acidic cation exchange resin. Journal of the
Chinese Institute of Chemical Engineers 30 (2), 117–122.

Lee, M.J., Wu, H.T., Lin, H.M., 2000. Kinetics of catalytic esterification

of acetic acid and amyl alcohol over Dowex. Industrial and Engineering
Chemistry Research 39, 4094–4099.

Lee, M.J., Chiu, J.Y., Lin, H.M., 2002. Kinetics of catalytic esterification of

propionic acid and n-butanol over Amberlyst 35. Industrial and Engineering
Chemistry Research 41, 2882–2887.

Lilja, J., Murzin, D.Yu., Salmi, T., Aumo, J., Mäki-Arvela, P., Sundell, M.,

2002. Esterification of different acids over heterogeneous and homogeneous
catalysts and correlation with the Taft equation. Journal of Molecular
Catalysis 182–183, 555–563.

Liu, W.T., Tan, C.S., 2001. Liquid-phase esterification of propionic acid with

n-butanol. Industrial and Engineering Chemistry Research 40, 3281–3286.

Liu, Y., Lotero, E., Goodwin Jr., J.G., 2006. Effect of water on sulfuric acid

catalyzed esterification. Journal of Molecular Catalysis A: Chemical 245,
132–140.

McKetta, J.J., 1983. Encyclopedia of Chemical Processing and Design, vol.

19. Marcel Dekker, New York, pp. 381–402.

Nagaraju, N., Mehboob, P., 1996. A comparison of catalytic activity of

zeolites with some Lewis acids in esterification reaction. Industrial Journal
of Chemical Technology 3, 253–255.

Namba, S., Wakushima, Y., Shimizu, T., Masumoto, H., Yashima, T., 1985.

Catalytic application of hydrophobic properties of high-silica zeolites.
Catalysis by Acids and Bases, 205–211.

Othmer, K., 1994. Encyclopedia of Chemical Technology, fourth ed. vol. 9.

Wiley, New York, pp. 781–811.

Pääkkönen, P.K., Krause, A.O.I., 2003. Diffusion and chemical reaction in iso-

amylene etherification within a cation exchange resin. Journal of Applied
Catalysis 245, 289–301.

Poling, B.E., Prausnitz, J.M., O’Connell, J.P., 2001. The Properties of Gases

and Liquids. fifth ed. McGraw-Hill, New York. pp. 8.75–8.111.

Pöpken, T., Götze, L., Gmehling, J., 2000. Reaction kinetics and chemical

equilibrium of homogeneously and heterogeneously catalyzed acetic acid
esterification with methanol and methyl acetate hydrolysis. Industrial and
Engineering Chemistry Research 39, 2601–2611.

Rao, B.A., Raghavaiah, C.V., Reddy, M.S., Chiranjivi, C., 1979. Esterification

of 1-butanol with propionic acid in a stirred tank batch reactor. Chemical
and Petrochemicals Journal, 7–12.

Rao, S.K.D.R., Sinha, R., Rao, R.J., 1978. Liquid phase catalytic esterification

of isopropanol with propionic acid. Chemical and Petrochemicals Journal,
11–14.

Rao, Y.V., Reddy, M.S., Rao, C.V., 1976. A study on the reaction kinetics of

n-octyl alcohol esterification with acetic acid in the presence of aluminum
sulphate catalyst. Chemical and Petrochemicals Journal, 27–35.

Rehfinger, A., Hoffmann, U., 1990. Kinetics of methyl tertiary butyl ether

liquid phase synthesis catalyzed by ion exchange resin-I. Intrinsic rate
expression in liquid phase activities. Chemical Engineering Science 45,
1605–1617.

Roy, R., Bhatia, S., 1987. Kinetics of esterification of benzyl alcohol with

acetic acid catalyzed by cation exchange resin (Amberlyst-15). Journal of
Chemical Technology and Biotechnology 37, 1–10.

Sai, P.S.T., 1988. Esterification kinetics of acetic acid with 1-butanol. Journal

of Energy Heat Mass Transfer 10 (2), 181–189.

Sharma, O.N., Nageshwar, G.D., Mene, P.S., 1973. Esterification of ethanol

with propionic acid catalysed by a cation exchange resin. Indian Journal
of Technology 11, 360–362.

Song, W., Venimadhavan, G., Manning, J.M., Malone, M.F., Doherty, M.F.,

1998. Measurements of residue curve maps and heterogeneous kinetics in
methyl acetate synthesis. Industrial and Engineering Chemistry Research
37, 1917–1928.

Venkateswarlu, K., Sinha, R., Rao, R.J., 1976. Kinetics of esterification of

n-butanol with propionic acid in a stirred tank batch reactor. Chemical
and Petrochemical Journal, 3–10.

Xu, Z.P., Chuang, K.T., 1996. Kinetics of acetic acid esterification over ion

exchange catalysts. Canadian Journal Chemical Engineering 74, 493–500.

Yadav, G.D., Kulkarni, H.B., 2000. Ion-exchange resin catalysis in the

synthesis of isopropyl lactate. Reactive and Functional Polymers 44,
153–165.

Yadav, G.D., Thathagar, M.B., 2002. Esterification of maleic acid with ethanol

over cation exchange resin catalysts. Reactive and Functional Polymers
52, 99–110.

Zhang, H.B., Zhang, K., Yuan, Z.Y., Zhao, W., Li, H.X., 1998. Esterification

reaction of acetic acid with n-amyl alcohol over zeolite beta. Journal of
Natural Gas Chemistry 7 (4), 336–345.


Document Outline


Wyszukiwarka

Podobne podstrony:
1 s2 0 S000925099800520X main
1 s2 0 S0009279709005614 main
1 s2 0 S0009254115002983 mainid Nieznany
1 s2 0 S0020025512001946 main
1 s2 0 S0378382002000085 main
1 s2 0 S0304397599001000 main
1 s2 0 S0006291X05021595 main
1 s2 0 S0040603111000104 main 2
1 s2 0 S0944501312001358 main
1 s2 0 S0166218X96000583 main
1 s2 0 S0005273614000303 main
1 s2 0 S0304397502001342 main
1 s2 0 S0377221798003622 main (1)
1 s2 0 S0022169496031423 main
1 s2 0 S1046592814002101 main
1 s2 0 0166218X93E0153P main
1 s2 0 S0022000006001474 main

więcej podobnych podstron