You are on page 1of 11

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright

Author's personal copy


Chemical Engineering Science 69 (2012) 201210

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Kinetic and diffusion study of acid-catalyzed liquid-phase alkyl


formates hydrolysis
a a, J.-P. Mikkola a,b
Olatunde Jogunola a,n, Tapio Salmi a, Johan Warn
a
b

Akademi, Process Chemistry Centre, Industrial Chemistry and Reaction Engineering, FI-20500 Abo/Turku,

Abo
Finland
Ume
a University, Technical Chemistry, ChemicalBiological Centre, Department of Chemistry, SE-90187 Ume
a, Sweden

a r t i c l e i n f o

abstract

Article history:
Received 8 December 2010
Received in revised form
3 October 2011
Accepted 11 October 2011
Available online 19 October 2011

Alkyl formates are typically hydrolyzed with the aid of homogeneous catalysts, but heterogeneous catalysts
provide a promising pathway for the process. The hydrolysis of alkyl formates by a solid acid catalyst, ionexchange resin was accomplished in a batch reactor, a stirred autoclave operating isothermally at 60 1C and
90 1C with a constant initial water-to-ester molar ratio. A mathematical model, which incorporates the
particle size distribution of the solid catalyst, was developed to study the kinetics and internal mass
transfer effects in the porous particles and it was able to predict the concentrations in the bulk phase and
inside the catalyst particles. A combined reactiondiffusion model is necessary to describe the behavior of
the system. The model was able to predict well the experimental results.
& 2011 Elsevier Ltd. All rights reserved.

Keywords:
Alkyl formate hydrolysis
Kinetics
Particle size distribution
Mathematical model
Internal diffusion limitation
Amberlite IR-120

1. Introduction
The consumption of formic acid, a medium volume commodity
chemical, has steadily increased since the European Unions ban
on the use of over-the-counter (OTC) feed antibiotics in livestock
feeds. Formic acid can be produced by hydrolysis of an alkyl
formate. The unfavorable equilibrium is a challenge for the process;
typically the equilibrium constant of alkyl formate hydrolysis has a
value less than 0.2. A trivial way to overcome this problem is to use
an excess of water to push the equilibrium to the side of the reaction
products. A process patented by BASF employed a weak base (an
amine), which generates a not too strong complex with the formic
acid product (Hohenschutz et al., 1979) thus giving a more favorable
equilibrium composition.
The process requires a catalyst too, since the reaction is slow in
neutral media. The catalyst can either be a base or an acid.
Mineral acids, such as sulfuric acid (Eversole, 1939) and hydrochloric acid (Long and Paul, 1957; Bell et al., 1955; Newling and
Hinshelwood, 1936), as well as organic acids, such as formic acid
(Leonard, 1979; Lynn et al., 1975) have been demonstrated to
catalyze the liquid-phase hydrolysis of alkyl formate. However,
only a few solid acid catalysts have been reported to be used for
the hydrolysis process because of the hydrophilic nature of water

Corresponding author. Tel.: 358 2 2154431; fax: 358 2 2154479.


E-mail address: jolatund@abo. (O. Jogunola).

0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.10.028

towards the catalyst. Among them are zeolites (Namba et al.,


1981), heteropolyacid catalysts (Izumi, 1997), modied zirconium
phosphonates (Segawa et al., 1993) as well as cation exchange resins
(Thomas and Davies, 1952; Haskell and Hammett, 1949; Metwally
et al., 1990; Ehteshami et al., 2006).
In the open literature, activated charcoal (Cho et al., 1980;
Wetherold et al., 1974) and acidic ion exchange resins (Mai et al.,
2004; Falk and Seidel-Morgenstern, 2002) have been successfully
used to catalyze the alkyl formate hydrolysis. This was achieved
with the aid of a chromatographic reactor in order to separate
different components and achieve a higher reactant conversion,
because the hydrolysis process is an equilibrium-limited reaction.
Ion exchange resins are cross-linked, insoluble polymers carrying
reversible xed functional groups or sites (Zagorodni, 2007).
These catalysts are often used at temperatures of 100 1C or less,
because a higher temperature leads to thermal instability of the
catalyst. Cation exchange resins, such as Amberlite series (Thomas
and Davies, 1952; Haskell and Hammett, 1949; Altiokka and C
- itak,

et al., 2000; Xu and Chuang,


2003) and the Amberlyst series (Popken
1997; Liu and Tan, 2001; Mazzotti et al., 1997; Shimizu and Hirai,
1986; Saha and Sharma, 1996) have been successfully used for
esterication and hydrolysis of alkyl acetates, but only very few
researchers have investigated alkyl formate hydrolysis on the resins.
Metwally et al. (1993) studied the hydrolysis of ethyl formate using
cation resins in an acetonewater mixture. Smith and Amundson
(1951) investigated the effect of intraparticle diffusion on ethyl
formate hydrolysis using Dowex 50 as a catalyst. They considered

Author's personal copy


202

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

the process as a simple reversible reaction, derived some equations


for three types of reactors and performed experiments on different
sizes of particles for continuous ow and batch reactors to verify the
equations. However, there has not yet been any detailed publication
concerning the kinetics of both ethyl and methyl formate hydrolyzes
in a batch reactor using cation resins.
Although, gel-type resins are not porous in the true sense since
their structure depends on swelling in the liquid in which they are
immersed. Thus, in their swollen state, their porosity increases
signicantly and we can identify an internal porosity in terms of
equilibrium uptake of the liquid (LeVan et al., 1997). Three different
intra-particle diffusion mechanism can be used to describe gel-type
exchanger namely pore diffusion, solid or surface diffusion and
parallel (both pore and surface) diffusion. The surface diffusion
model is frequently used to describe mass transfer in gel-type resin;
however, pore diffusion model may be used to evaluate the internal
porosity available to the liquid since in their swollen state, they have
an appreciable porosity. Furthermore, pore diffusion model is very
useful in describing the catalyst effectiveness factor.
This work examines the kinetic modeling of alkyl formate
hydrolysis on Amberlite IR-120 resin. The particle size distribution of the solid catalyst is incorporated in the model to reveal the
kinetics and the internal mass transfer limitation encountered in
the solid acid catalyst.

2. Experimental section
2.1. Experimental set-up, procedure and matrix
The reactor is a conventional 500 ml Parr autoclave made of
zirconium metal. It consists of feeding and reaction vessels, a
heating unit, a sampling line and a stirrer. The sampling line
connected to a cooling bath is equipped with a lter. Two
thermocouples and cooling air are used for controlling temperature. The hydrolysis reaction was performed isothermally at a
nitrogen-pressure of 20 bar and at temperatures of 60 1C and
90 1C using Amberlite IR-120 (Aldrich) as a catalyst.
The following procedure was applied in the kinetic experiments. Distilled water with the catalyst was placed in the reaction
vessel. Ethanol (Altia, 99.5%) and ethyl formate (Sigma-Aldrich,
97 wt%) in case of ethyl formate hydrolysis (EFH) or methanol
(Baker, 499 wt%) and methyl formate (Sigma-Aldrich, 97 wt%) in
case of methyl formate hydrolysis (MFH) were discharged into a
feeding vessel with the aid of a dropping funnel. A nitrogen line to
the reactor and vent lines were connected to the system. The
stirred reaction vessel was heated until it reached the desired
temperature. The content in the feeding vessel was discharged
into the reaction vessel as quickly as possible using nitrogenpressure and the reaction commenced immediately. There was a
temperature drop in the initial stages and the temperature
increased gradually until it reached the desired temperature after
less than 5 min. The initial total amount of the hydrolysis mixture
was close to 0.3 kg. Table 1 summarizes the experimental details.

Table 1
Experimental matrix.
Reactants/catalyst

MFH

EFH

C0A (mol/kg)
[RFo]0
[ROH]0
[H2O]0
Catalyst loading (g/kg on dry basis)
Stirring rate

10.49
0.94
18.87
1.6716.67
300 rpm

9.04
0.44
16.28
3.3366.67

RCH3 for MFH (methyl formate hydrolysis) or CH3CH2 for EFH (ethyl formate
hydrolysis).

Table 2
Properties of Amberlite IR 120 resin.
Properties

Specication

Bead type
Cross linking (%)
Particle size range (mm)
Moisture content (% mass)
Density of wet catalyst (g/cm3)
Maximum operating temperature (1C)
Capacity by dry weight (meq/g)

Gel
8
0.31.2
45
1.26
120
4.4

time for the sample was between 6 and 11 min. The experimental
results were based on the alcohol analysis (ethanol in the case of
EFH and methanol in the case of MFH) due to the difculty in getting
a reliable analysis method for ethyl formate and formic acid and the
volatility of methyl formate.
2.3. Catalyst properties and characterization
The properties of the solid catalyst as given by the manufacturer are depicted in Table 2.
The catalyst was pretreated by washing with distilled water,
decanted and then dried in an oven at 99 1C for two days until a
constant mass is obtained. So, all experiments were performed
with the dried resins. The purpose of a dry resin is to know the
exact quantity of water in the system since water is one of the
reactants.
The particle size distribution of the dried resins was determined by the classical sieving analysis method and in water at
room temperature using a laser diffraction particle sizer, the
Malvern 2600 series. The working principle of the instrument is
based on the light scattering property of the particles using
HeNe laser diffraction system.
The acid sites capacity was measured by a conventional
titration method (Zagorodni, 2007), which has been described in
an earlier publication (Jogunola et al., 2010).

3. Catalyst characterization and effectiveness factor

2.2. Gas chromatographic analysis

3.1. Particle size distribution

Samples were taken in dened sample intervals (5, 10, 15 or


20 min) using the reactor system pressure and were analyzed offline with a gas chromatograph: 6890N, injection port temperature
150 1C, oven temperature 150 1C, HP-PLOT U column, 250 m 
530 mm  20 mm, carrier gas helium (15 ml/min at 1 min), detector
FID (250 1C, H2 ow 40 ml/min, air ow 450 ml/min). The calibration was performed using acetonitrile (Labscan, 499%) as an
internal standard. The calibration of the internal standard solution
was always done before the analysis of any sample. The analysis

A rough estimate of the particle size distribution of the dry


resin was achieved by using the classical sieving method. The
proportion of the particles with diameters less than 0.5 mm was
41.76 wt% and those of particles with diameter larger than
0.5 mm was 58.24 wt %. However, the particle size distribution
of the dry resin immersed in water after few seconds as obtained
with the Malvern 2600 series instrument is given in Fig. 1.
From the gure, the particle size range 0.060.08 mm constitutes 0.8 vol% of the distribution, while 0.120.26 makes up

Author's personal copy


203

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

18
16
14
Vol-%

12
10
8
6
4
2
0.97-1.13

0.84-0.97

0.72-0.84

0.63-0.72

0.54-0.63

0.47-0.54

0.40-0.47

0.35-0.40

0.30-0.35

0.26-0.30

0.22-0.26

0.19-0.22

0.17-0.19

0.14-0.17

0.12-0.14

0.11-0.12

0.09-0.11

0.08-0.09

0.07-0.08

0.06-0.07

Particle diameter range (mm)


Fig. 1. Particle size distribution of Amberlite IR-120 using Malvern 2600 series.

18
dp < 0.5 mm
14
wt-% EtOH

6.1 vol%. The distribution is continuous in the particle size range


0.31.13 mm, which constitutes 92.8 vol%. The mean diameter of the
particle according to the instrument is 0.53 mm. The specic surface
area according to the instrument is 0.0156 m2/cm3 (o0.1 m2/g).
However, this value is not as reliable as that using a surface area
specic technique e.g. argon or nitrogen adsorption porosimetry.
However, it was difcult to obtain the specic surface area using
nitrogen BET method because in the dry state, Amberlite IR 120 resin
has no effective permanent porosity.

0.3 < dp < 0.5 mm


10
dp > 0.5 mm
6

3.2. Effect of catalyst particle sizeeffectiveness factor

observed rate of ester disappearance


rate in the absence of internal diffusion

From the above equation, the effectiveness factor for particle


sizes 40.5 mm and 0.31.2 mm can be estimated. The results are
displayed in Table 3.
Thus, there is internal mass transfer limitation at the beginning of
the reaction for larger particles (dp 40.5 mm), but the effect
diminishes as the equilibrium is approached and the reaction rate is
slowed down (see Table 3 and Fig. 2). However, for MFH (Fig. 2b),
there was no observed signicant difference between the experimental curves using particles with dpo0.5 mm or 0.3odpo1.2 mm.
This might be due to experimental error. However, the observations
are reasonable; the mass transfer resistance plays the most dominant
role when the reaction is at its fastest. The presence of internal
diffusion limitation has been previously conrmed in an earlier
publication (Jogunola et al., 2010).

40
60
Time (min)

20

80

100

16

12
Wt-% MeOH

At a stirring speed range of 300700 rpm, there was no external


mass transfer limitation and attrition of the catalyst was not visible
under the chosen reaction condition. So, 300 rpm was chosen for the
entire experiments. The pretreated (dried catalyst) Amberlite IR120, with a particle size of 0.31.2 mm was sieved with a 0.5 mm
sieve and categorized into two different size ranges, i.e. o0.5 mm
and 40.5 mm. The effect of internal mass transfer resistance inside
the catalyst with the same catalyst loading of these two size ranges
and the pretreated resin without sieving was tested on the hydrolysis of both ethyl and methyl formate. The effect of the internal
mass transfer inside the particle for both EFH and MFH, respectively,
is illustrated in Fig. 2a and b, respectively.
The smallest particle (dp o0.5 mm) has the fastest kinetics.
Assuming this particle represents the intrinsic kinetics, the limiting rate, the effectiveness factor Z can be estimated as

dp < 0.5 mm
8

0.3 < dp < 1.2 mm

4
dp > 0.5 mm
0
0

20

40

60

Time (min)
Fig. 2. (a) Experimental determination of the effectiveness factor for EFH using
2.5 g catalyst at 60 1C, H2O/EtFo 1.8. (b) Experimental determination of the
effectiveness factor for MFH using 2.5 g catalyst at 60 1C, H2O/MeFo 1.8.

4. Kinetic and diffusion models


It is well known that an ion exchange resin absorbed liquid
and swelled when it comes in contact with that liquid. Thus, the
concentration of the reacting liquid inside the catalyst is different
from the bulk liquid. Consequently, the kinetics of hydrolysis of
esters or its reverse reaction, esterication on the resin has been
described by different models. Xu and Chuang (1997) used the

Author's personal copy


204

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

Table 3
Effectiveness factor for different size ranges. of the (dry) pretreated Amberlite
IR-120.

The equilibrium constant is related to the reaction


enthalpy through the vant Hoff equation, therefore
o
K eq K 0,eq eDHR =R1=T1=T ref .

Particle size (mm)

t 0 min

t 20 min

t 50 min

EFH
dp o0.5
0.3 odp o1.2
dp 40.5

1.00
0.92
0.88

1.00
0.91
0.78

1.00
0.95
0.89

MFH
dp o0.5
0.3 odp o1.2
dp 40.5

1.00
0.97
0.55

1.00
0.96
0.71

1.00
0.99
0.85

4.2. Heterogeneous catalysis


To describe the heterogeneously catalyzed process, a pseudohomogenous model was used, which did not take into account the
selective adsorption on the catalyst surface. The catalytic kinetics
of the reaction can thus be described by


CC CD
0
r 0i k C A C B 
4
K eq
where
E0

pseudo-homogenous model to describe esterication of acetic


acid with methanol. The adsorption-descriptive model comprising EleyRideal (Altiokka, 2007; Lilja et al., 2002) and Langmuir

Hinshelwood (Popken
et al., 2000; Song et al., 1998), where the
concentration of the adsorbed species were taken into consideration,
have also been used to describe both esterication and hydrolysis
reactions. Another model, which has been used to describe the
reaction rate of resin, is the absorption model (Mazzotti et al., 1997;
Sainio et al., 2004). Here, the phase equilibrium between the liquid
and the resin phase, alongside a reaction on the resin phase are
taken into account. Both absorption-and adsorption-descriptive
models are complicated and involve independent runs to evaluate
both absorption and adsorption equilibrium constants. Thus, a
simplied model, pseudo-homogenous, which ignored both phenomena, was used for our work. However, this model cannot be
used for the purpose of extrapolation (only concentration for which
it was developed).
A characteristic feature of the system (i.e. hydrolysis of alkyl
formate) is that the process takes places via simultaneous reaction
paths: a homogeneous catalysis induced by the reaction product,
formic acid as well as a heterogeneous catalysis on the surface of the
ion exchange resin. The latter reaction path is the dominating one,
but homogeneous catalysis cannot be completely excluded in the
kinetic treatment.
4.1. Homogenous catalysis
The reversible hydrolysis of alkyl formate in the liquid-phase
can be considered as follows:
A B$C D
where A is ethyl formate (EtFo) or methyl formate (MeFo), B is
water, C is formic acid and D is ethanol (EtOH) or methanol (MeOH).
The reaction rate in the absence of the catalyst is expressed as
r i f C A C B 2

1
CC CD
K eq

where f is the rate function, Keq is the concentration-based


equilibrium constant and C is the concentration. The rate function
is dened by
f k kauto C C

where k is the rate constant due to the dissociation of B, and kauto is


the rate constant of the autocatalytic reaction. Furthermore, using
the temperature dependence of the rate constants by Arrhenius
equation (k AeEA =RT ), the reaction rate can be described as follows:
E

ri

 RA

ke

1
1
T T ref

kauto C C e

EA,auto
R

1 1
T T ref

 !

CACB

CC CD
K eq

A
0 
k0 e R

1
1
T T ref

It should be noted that the rate constant (k0 ) in the above


expression incorporates the concentration of the active surface sites
on the catalyst, i.e., the acid groups on the ion exchanger (SO3H).
Thus, the intrinsic kinetics can be described by the following
rate equation:


CC CD
0
r tot k kauto C C k C A C B 
5
K eq
Keq is based on the liquid phase concentration, therefore an
ideal liquid mixture is assumed, where the liquid volume remains
constant. The following values were used for the reaction enthalpies; methyl formate DHoR 5.44 kJ/mol at 25 1C and ethyl
formate DHoR 7.91 kJ/mol at 25 1C (Roine, 2009).
4.3. Reactor and catalyst particle model with particle size
distribution
The simultaneous chemical reaction and diffusion in the
porous ion-exchange particles are described by a dynamic model,
coupled parabolic partial differential equations. The diffusion mathematical treatment is categorized into catalyst particle model and
the batch reactor with particle size distribution. The mass balance
for the catalyst particle is
!
2
r 0 rp
dC pi
d C pi a1 dC pi
D
i r i ei2
6

dt
ep
X dX
ep Rj dX 2
The boundary conditions are Cpi(X 1) Ci at the catalyst particle surface and dC pi =dXX 0 0 at the particle center. Because of
vigorous stirring (Table 1), the external mass transfer effect was
negligible. The catalyst particle mass balance (reactiondiffusion
model) was solved numerically for each particle size fraction to
obtain the concentration proles in the particles. The effective
diffusion coefcient were obtained from the relation Dei eP =tP Di ,
while the molecular diffusion coefcient of each component was
calculated with the aid of the WilkeChang equation (Reid et al.,
1988)
p
7:4  1012 yM T 2
Di
m =s
7
0:6
mm V i
valid for the random pore model. An average porosity was applied for
the particle. The viscosity of water at the desired temperature was
used for the mixture viscosity and the molar volumes of the
components, Vi were obtained from atomic contribution of LeBas.
A batch reactor model incorporating the experimentally determined particle size distribution of the catalyst was used in this
work. Detailed derivation of the model is given by Leveneur
et al. (2009). For the bulk phase, the component mass balance is

Author's personal copy


205

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

written as

Table 4
Estimated parameters of EFH in the absence of the catalyst.
Estimated Parameter
parameter value
Keq
EA
k
kauto
EAauto

0.107
74.5 kJ/mol
3.12  10  5
L/mol min
1.61  10  4
L2/mol2 min
58.9 kJ/mol

Relative
error (%)

Correlation matrix
EA
k
Keq

2.3
10.8
12.1

1.00
0.038
0.310

18.5

 0.940  0.524  0.370 1.00

17.8

 0.533  0.975  0.076 0.658 1.00

1.00
0.017

kauto

X
dC i
ap
yj N ij x2j r i
dt
j

EAauto

1.00

Tref 75 1C (average temperature); degree of explanation 99.62%, EFH ethyl


formate hydrolysis.

Table 5
Estimated parameters of MFH in the absence of the catalyst.
Estimated Parameter
parameter value
Keq
EA
k
kauto
EAauto

0.111
64.4 kJ/mol
1.68  10  5
L/mol min
1.66  10  4
L2/mol2 min
47.1 kJ/mol

Relative
error (%)

Correlation matrix
Keq
EA
k

2.4
19.8
19.1

1.00
 0.041 1.00
0.228
0.465

26.6

 0.248  0.460  0.997 1.00

30.2

0.035

kauto

EAauto

1.00

where ap is the total outer particle surface area-to-liquid volume,


yj is the fraction of particles with the radius Rj and xj is the
normalized particle radius (Rj/Rav, see Section Nomenclature). The
uxes Nij at outer surface of the catalyst particle are obtained by
integration of the proles inside the catalyst particles.
The reactiondiffusion model (Eqs. (6) and (8)) were solved
numerically by discretizing the partial differential equations
(PDEs) with respect to the spatial co-ordinate (X). Central nite
difference formulae were used to approximate the rst and
second derivatives. Thus, the PDEs were transformed to ordinary
differential equations (ODEs), which were solved numerically
with the backward difference method.
For the solution of the model equations and for parameter
estimation purposes, the software Modest (Haario, 2007) was
used. The software solves the model equations and minimizes
the objective function Q S(Ci,model  Ci,exp)2 by adjusting the
kinetic parameters with the LevenbergMarquardt or simplex
method.

5. Kinetic modeling results and discussion

 0.995  0.487 0.485 1.00

Tref 75 1C (average temperature); degree of explanation 99.4%, MFH methyl


formate hydrolysis.

During the analysis of the samples, no side reaction (i.e.


selectivity is 100%) was observed for all experiments. The results
were based on the analysis of the alcohol formed since alkyl
formate analysis gave inconsistent results. Homogenous catalysis

60 C

1.2

90 C

1.4
1.2

mol

mol

0.8
0.6

0.8
0.6

0.4

0.4

0.2

0.2
0

0
0

100

200
300
400
time (min)

500

60 C

1.4

50

100
time (min)

150

200

150

200

90 C

1.5

1.2
1

0.8

mol

mol

0.6

0.5

0.4
0.2

0
0

100

200 300 400


time (min)

500

600

50

100
time (min)

Fig. 3. Fit of the model to experimental data (alcohol) in the absence of solid catalyst for (a) ethyl formate hydrolysis and (b) methyl formate hydrolysis; H2O/ester 1.8.

Author's personal copy


206

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

played a signicant role in the rate expressions since the reaction


is autocatalyzed in the absence of solid acid catalyst (Eq. (3)).
In addition, experimental analysis of the system showed that

internal diffusion in the catalyst particles plays a dominant role.


The kinetic parameters for the non-catalytic reaction were rst
estimated using experimental data obtained in the absence of

Table 6
Estimated parameter values for the catalytic reaction in the presence of diffusion limitation.

Ethyl formate
Explained 98%
Methyl formate
Explained 98%

Estimated
parameter

Parameter
value

Relative
error (%)

Correlation matrix
ep/tp
k0

E0 A

ep/tp
k0
E0A
ep/tp
k0
E0A

0.018
1.8  10  4 mol/gcat min
32.2 kJ/mol
0.018 xed value
1.0  10  3 mol/gcat min
50.8 kJ/mol

18.3
26.9
190

1.00
 0.194
0.943

1.00

3.7
5.3

1.00
 0.218
1.00
0.308

1.00

Tref 75 1C (average temperature).

1g

2.5 g

1
mol

1.5

mol

1.5

0.5

0.5

1.5

100
200
time (min)
5g

300

50

100
time (min)

150

20 g
1.5

1
mol

mol

0.5

0.5

0
0

50
time (min)

100

1g

100
time (min)

150

2.5 g

mol

1.5

mol

1.5

50

0.5
0

0.5

50

100
time (min)

150

5g

1.5

100

10 g

1.5
1
mol

mol

50
time (min)

0.5
0

0.5

50
time (min)

100

50
time (min)

100

Fig. 4. Fit of the model to experimental data (ethanol) for ethyl formate hydrolysis using different amounts of the catalyst and at (a) 60 1C and (b) 90 1C; H2O/ester 1.8.

Author's personal copy


207

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

the catalyst, while keeping DHor value constant. The estimated


parameter values for the non-catalytic reaction are displayed in
Tables 4 and 5 for ethyl formate and methyl formate hydrolysis,
respectively.
The model t to the experimental data of both ethyl and
methyl formate hydrolysis for the non-catalyzed reaction is
depicted in Fig. 3.
The values of the parameters obtained from the non-catalytic
reaction were used as input values to estimate the heterogeneous
catalysis (Eq. (5)). For the catalytic reaction, both reaction and
diffusion in the catalyst particles were taken into consideration.
The catalyst mass balances were solved simultaneous for 12
particle sizes with different fractions and the molecular diffusion
coefcients of the components of the hydrolysis solution were
obtained from WilkeChang equation. The viscosity of water was
used in the calculation. This information was incorporated in the
model; in which the effective diffusion coefcients play a central
role. The particle porosity was xed at 0.5 and the porosity-to-

tortuosity ratio in the diffusion coefcients was adjusted in situ


during the simulations using non-linear regression analysis.
For the methyl formate reaction, the porosity-to-tortuosity
ratio was xed to the same value as obtained from the ethyl
formate reaction. This is justied, since the catalyst was the same
in both cases. The estimation of the porosity-to-tortuosity ratio
for the methyl formate reaction led to a heavy correlation
between the rate constant and the porosity-to-tortuosity ratio.
The parameter values for the catalytic reaction for both ethyl
and methyl formate hydrolyzes are collected in Table 6.
From the table, it can be deduced that methyl formate
hydrolysis is faster than ethyl formate hydrolysis under similar
experimental conditions. This can be attributed to the more steric
hindrance present in ethyl formate molecule. The activation
energy of resin-catalyzed methyl formate hydrolysis is consistent
with that reported in the literature (Metwally et al., 1993).
However that of ethyl formate is lower that the literature value.
This is exemplied in the relative error. This might be due to the

1g

2.5 g
1.5

mol

mol

1.5

0.5

0.5

0
0

50
100
time (min)
5g

150

50
time (min)

100

50
time (min)

100

1.5

mol

1
0.5
0

0.5 g

1g
1.5

1
mol

mol

1.5

0.5
0

0.5

50
time (min)
2g

100

50
time (min)

100

1.5

50
time (min)

100

mol

1
0.5
0

Fig. 5. Fit of the model to the experimental data (methanol) for methyl formate hydrolysis at (a) 60 1C and (b) 90 1C using different amounts of the solid catalyst; H2O/
ester 1.8.

Author's personal copy


208

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

10.5
10
9.5
9
C mol/dm3

fact that selective sorption, the swelling of the resin and the nonideality of the system were not taken into consideration in the
treatment of the model. Furthermore, the tortuosity value is
unexpectedly high and not in the correct range. This implies that
the pore path is more sinuous and the pore has a complex
connectivity and constrictivity. Therefore, a more detailed evaluation of the system is necessary. Figs. 4 and 5 show the t of the
model to the experimental data for both EFH and MFH (mol%
alcohol), respectively.
In general, we can conclude that the model t is very good, as
revealed by Figs. 35: the predicted concentrations follow the
experimental ones very closely and no systematic deviations
appear. The correct equilibrium state is always approached by
the model. The degree of explanation is always high (Tables 46),
which reects the very good overall t of the model. The relative
errors of the parameters and the mutual correlations between the
parameters are on a reasonable level, taking the complexity of the
model (the many parameters included) into account.
Fig. 3 reveals the interesting autocatalytic effect: in the beginning, the reaction rate is very slow, but it is enhanced as soon as
some formic acid is released; formic acid then gives a very clear
autocatalytic contribution as predicted by the rate equation (Eq. (3)).

8.5
8
7.5
0.15
0.18 0.21

7
6.5

0.440.67
0.37
0.58 1.1
0.32
0.5 0.9
0.79

6
0

0.1

0.2

0.3

0.4

0.5
x

0.6

0.7

0.8

0.9

9.5
9
8.5
c mol/dm3

8.5
8

c mol/dm3

7.5

7.5
7

7
0.15

6.5

0.18

0.1

0.2

0.3

0.4

0.5
x

0.6

0.37 0.44

0.67
0.58
0.91.1
0.79

0.5

5.5

0.21 0.32
0.79
0.370.5
0.581.1

0.21
0.32

0.15
0.18

6.5

0.7

0.8

0.9

0.1

0.2

0.3

0.4

0.5
x

0.6

0.7

0.8

0.9

Fig. 7. Concentration proles of methyl formate hydrolysis inside the catalyst


particle at (a) 60 1C using 2.5 g catalyst and (b) 90 1C using 2 g for different particle
sizes (particle diameter in mm) in the beginning of the reaction; H2O/ester 1.8.

Table 7
The molecular diffusion coefcient of the components of
the hydrolysis mixture.

Components

7.5
7
c mol/dm3

6.5
0.15

0.18
0.21

0.37
0.32

5.5

0.44

1.1

0.58
0.5

0.67

0.1

0.2

0.3

0.4

0.5
x

0.6

0.7

Ethyl formate hydrolysis


Ethyl formate
Water
Formic acid
Ethanol

2.60
6.18
4.11
3.10

Methyl formate hydrolysis


Methyl formate
Water
Formic acid
Methanol

3.20
6.60
4.39
4.39

0.9
0.79

5
0

Diffusion coefcient
Di (10  9 m2/s) at 60 1C

0.8

0.9

Fig. 6. Concentration proles of ethyl formate hydrolysis inside the catalyst


particle at (a) 60 1C and (b) 90 1C for different particle sizes (particle diameter in
mm) in the beginning of the reaction using 5 g catalyst; H2O/ester 1.8.

Some simulated concentration proles for ethyl and methyl


formate hydrolysis inside the catalyst particle are shown in
Figs. 6 and 7, respectively. The gures demonstrate the presence
of the internal diffusion limitation in the system. Methyl and
ethyl formate behave in a very similar manner, which is expected
from the rather similar diffusion coefcients (Table 7) for both

Author's personal copy


O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

systems. The diffusion limitation illustrated in the gures represents


the situation in the beginning of the reaction; as the reaction
proceeds, the rate is retarded and the diffusion effect diminishes.
The diffusion effect is most prominent for the largest particles, but
even for the smallest particles, some diffusion effect can be observed.

6. Conclusions
Extensive experiments were carried out for the solid-catalyzed
hydrolysis of methyl and ethyl formates in a well-stirred, isothermal batch reactor. A combined kinetics-diffusion model was
developed for the system, comprising both homogeneous and
heterogeneously catalyzed kinetics as well as internal diffusion
effects. The model parameters were estimated with non-linear
regression analysis. The model t was very good and the presence
of coupled kinetic-internal diffusion effects was clearly demonstrated by the model. However, the tortuosity value was unexpectedly high. Even though the model was used for data obtained
from a batch reactor, it is expected that the same effects prevail in
continuous xed bed reactors, too. Thus, the approach is not
limited to batchwise operated systems only, but it can in future
be used for continuous systems for alkyl formate hydrolysis.

Nomenclature
A
a
ap
C
D
De
dp
EA
f
DH
K
k
M
N
Q
R
Rj
ri
r
r
T
t
V
X
xj
yj

ep
y

m
rp
tp

pre-exponential factor
shape factor of a catalyst particle (a 3 for spheres)
outer catalyst particle area-to-liquid volume
concentration
diffusion coefcient
effective diffusion coefcient
particle diameter
activation energy
rate function
reaction enthalpy
equilibrium constant
rate constant
molar mass
diffusion ux
objective function
general gas constant
particle radius
generation rate for component i
radial coordinate
reaction rate
temperature
time
molar volume at normal boiling point
dimensionless particle coordinate (Xr/Rj)
normalized particle radius (xj Rj/Raver)
fraction of particles with the radius rj
porosity
association factor
dynamic viscosity
density of a catalyst particle
tortuosity of a catalyst particle

Subscripts and superscripts


auto
aver
cat
eq
exp

autocatalysis
average value
catalyst
equilibrium
experimental value

i
j
m
model
p
ref
S
tot
0
0

209

component index
particle size index
mixture
value predicted by the model
particle
reference
solvent
total
initial
heterogeneous-catalyzed reaction

Abbreviations
EFH
ethyl formate hydrolysis
FA (C)
formic acid
H2O (B) water
EtFo (A), MeFo (A) ethyl and methyl formate
EtOH (D), MeOH (D) ethanol, methanol
MFH
methyl formate hydrolysis

Acknowledgments

This work is a part of the activities at the Abo


Akademi Process
Chemistry Centre (PCC) within the Finnish Centre of Excellence
Programmes (20002005 and 20062011) by the Academy of

Finland. Financial support from the Abo


Akademi Foundation is
gratefully acknowledged.
References
Altiokka, M.R., 2007. Kinetics of hydrolysis of benzaldehyde dimethyl acetal over
Amberlite IR 120. Ind. Eng. Chem. Res. 46, 10581062.
Altiokka, M.R., C
- itak, A., 2003. Kinetics study of esterication of acetic acid with
isobutanol in the presence of amberlite catalyst. Appl. Catal. A: Gen. 239, 141148.
Bell, R.P., Dowding, A.L., Noble, J.A., 1955. The kinetics of ester hydrolysis in
concentrated aqueous acids. J. Chem. Soc. 9, 31063110.
Cho, B.K., Carr, R.W., Aris, R., 1980. A continuous chromatographic reactor. Chem.
Eng. Sci. 35, 7481.
Ehteshami, M., Rahimi, N., Eftekhari, A.A., Nasr, M.J., 2006. Kinetic study of
catalytic hydrolysis reaction of methyl acetate to acetic acid and methanol.
Iran. J. Sci. Tech. Trans. B Eng. 30, 595606.
Eversole, J.F., May 30 1939. Manufacture of Formic Acid. US Patent 2160064, to
Carbide and Carbon Chemicals.
Falk, T., Seidel-Morgenstern, A., 2002. Analysis of discontinuously operated
chromatographic reactor. Chem. Eng. Sci. 57, 15991606.
Haario, H., 2007. ModEst 6A Users Guide. Profmath Oy, Helsinki.
Haskell, V.C., Hammett, L.P., 1949. Rates and temperature coefcients in the
hydrolysis of some aliphatic esters with a cation exchange resin as the
catalyst. J. Am. Chem. Soc. 71, 12841288.
Hohenschutz, H., Schmidt, J.E., Kiefer, H., 1979. Process for Manufacturing Formic Acid.
Eur. Patent 1432 to BASF, April 18 1979, (translated from German to English).
Izumi, Y., 1997. Hydration/hydrolysis by solid acids. Catal. Today 33, 371409.

Jogunola, O., Salmi, T., Eranen,


K., Mikkola, J.-P., 2010. Qualitative treatment of
catalytic hydrolysis of alkyl formates. Appl. Catal. A Gen. 384, 3644.
Leonard, J.B., 1979. Preparation of Formic Acid by Hydrolysis of Methyl Formate.
Eur. Patent 5998, December 12 1979.
LeVan, M.D., Carta, G., Yon, C.M., 1997. Adsorption and ion exchange. in: Green,
D.W. (Ed.), Perrys Chemical Engineers HandbookSeventh ed. , McGraw-Hill,
New York.
a,
J., Salmi, T., Murzin, D., Estel, L., 2009. Interaction of intrinsic
Leveneur, S., Warn
kinetics and internal mass transfer in porous ion-exchange catalysts. Chem.
Eng. Sci. 64, 41014114.

Lilja, J., Murzin, D.Yu., salmi, T., Aumo, J., Maki-Arvela,


P., Sundell, M., 2002.
Esterication of different acids over heterogeneous and homogeneous catalysts and correlation with the Taft equation. J. Mol. Catal. A 182183, 555563.
Liu, W.-T., Tan, C.-S., 2001. Liquid-phase esterication of propionic acid with
n-butanol. Ind. Eng. Chem. Res. 40, 32813286.
Long, F.A., Paul, M.A., 1957. Application of the H0 acidity function to kinetics and
mechanism of acid catalysis. Chem. Rev. 57, 9351010.
Lynn, J.B., Homberg, O.A., Singleton, A.H., 1975. Formic Acid Synthesis by Lower
Alkyl Formate Hydrolysis. US Patent 3907884, September 23 1975, to
Bethlehem Steel Corp.
Mai, P.T., Vu, T.D., Mai, K.X., Seidel-Morgenstern, A., 2004. Analysis of heterogeneously catalyzed ester hydrolysis performed in a chromatographic reactor
and in a reaction calorimeter. Ind. Eng. Chem. Res. 43, 46914702.

Author's personal copy


210

O. Jogunola et al. / Chemical Engineering Science 69 (2012) 201210

Mazzotti, M., Neri, B., Gelosa, D., Kruglov, A., Morbidelli, M., 1997. Kinetics of
liquid-phase esterication catalyzed by acidic resins. Ind. Eng. Chem. Res. 36,
310.
Metwally, M.S., Razik, A.A., El-Hadi, M.F., El-Wardany, M.A., 1993. Rates and
temperature coefcients in resin-catalyzed hydrolysis of aliphatic esters.
React. Kinet. Catal. Lett. 49, 151159.
Metwally, M.S., El-Hadi, M.F., El-Wardany, M.A., Razik, A.A., 1990. Synthesis of a
new sulphonated cation exchange resin and its application in catalysed
hydrolysis of esters. J. Mater. Sci. 25, 42234225.
Namba, S., Hosonuma, N., Yashima, T., 1981. Catalytic application of hydrophopic
properties of high-silica zeolite. J. Catal. 72, 1620.
Newling, W.B.S., Hinshelwood, C.N., 1936. The kinetics of the acid and the alkaline
hydrolysis of esters. J. Chem. Soc., 13571361.

Popken,
T., Gotze,
L., Gmehling, J., 2000. Reaction kinetics and chemical equilibrium
of homogeneously and heterogeneously catalyzed acetic acid esterication with
methanol and methyl acetate hydrolysis. Ind. Eng. Chem. Res. 39, 26012611.
Reid, R.C., Prausnitz, J.M., Poling, B.E., 1988. The Properties of Gases and Liquids,
Fourth ed. McGraw-Hill, New York.
Roine, A., 2009. HSc Chemistry 7.0. Outokumpu Research Oy, Pori.
Saha, B., Sharma, M.M., 1996. Esterication of formic acid, acrylic acid and
methacrylic acid with cyclohexene in batch and distillation column reactors:
ion-exchange resins as catalysts. React. Funct. Polym. 28, 263278.

Sainio, T., Laatikainen, M., Paatero, E., 2004. Phase equilibria in solvent mixture-ion
exchange resin catalyst systems. Fluid Phase Equilib. 218, 269283.
Segawa, K., Kihara, N., Yamamoto, H., Nakata, S., 1993. Molecular design of layered
zirconium phosphate catalysts. in: Ballmoos, et al. (Eds.), Proceedings of the
Ninth International Zeolite Conference, Butterworth-Heinemann, Montreal,
pp. 255262.
Shimizu, S., Hirai, C., 1986. Catalysis by microreticular sulfonic acid resin in
esterication. Bull. Chem. Soc. Jpn. 59, 711.
Smith, N.L., Amundson, N.R., 1951. Intraparticle diffusion in catalytic heterogeneous systems. Ind. Eng. Chem. 43, 21582167.
Song, W., Venimadhavan, G., Manning, J.M., Malone, M.F., Doherty, M.F., 1998.
Measurement of residue curve maps and heterogeneous kinetics in methyl
acetate synthesis. Ind. Eng. Chem. Res. 37, 19171928.
Thomas, G.G., Davies, C.W., 1952. Ion-exchange resins as catalysts in hydrolysis of
esters. J. Chem. Soc. 295, 16071610.
Wetherold, R.G., Wissler, E.H., Bischoff, K.B., 1974. An experimental and computational study of the hydrolysis of methyl formate in a chromatographic reactor.
Adv. Chem. Ser. 133, 181190.
Xu, Z.P., Chuang, K.T., 1997. Effect of internal diffusion on heterogeneous catalytic
esterication of acetic acid. Chem. Eng. Sci. 52, 30113017.
Zagorodni, A.A., 2007. Ion Exchange Materials Properties and Applications, Elsevier,
Amsterdam.

You might also like