Sunteți pe pagina 1din 10

Chemical Engineering Journal 264 (2015) 258267

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Thermodynamic and kinetic studies for synthesis of the acetal


(1,1-diethoxybutane) catalyzed by Amberlyst 47 ion-exchange resin
Mehabub Rahaman, Nuno S. Graa, Carla S.M. Pereira, Alrio E. Rodrigues
Laboratory of Separation and Reaction Engineering (LSRE), Department of Chemical Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200-465
Porto, Portugal

h i g h l i g h t s
 Synthesis of 1,1-diethoxybutane was studied in a batch reactor.
 Thermodynamic equilibrium constant was determined in the temperature range of 293.15323.15 K.
 The standard formation properties of 1,1-diethoxybutane have been determined.
 Two-parameter kinetic model based on a LangmuirHinshelwood indicates good agreement with the experimental data.

a r t i c l e

i n f o

Article history:
Received 5 June 2014
Received in revised form 14 November 2014
Accepted 15 November 2014
Available online 25 November 2014
Keywords:
Synthesis of 1,1-diethoxybutane
Batch reactor
Amberlyst 47 catalyst
Kinetics
Thermodynamics

a b s t r a c t
The synthesis of 1,1-diethoxybutane was carried out in a batch reactor from a liquid phase reaction
between ethanol and butyraldehyde using Amberlyst 47 as the solid acid catalyst to obtain thermodynamic, kinetic and adsorption parameters. The reaction equilibrium constant was experimentally determined in the temperature range 293.15323.15 K. The standard properties of the reaction at 298.15 K
were also estimated. The effects of temperature, molar ratio of ethanol to butyraldehyde, stirrer speed,
and catalyst loading on the reaction rate were investigated. Kinetic experiments were performed in
the same temperature range at 6 bar. A two-parameter kinetic law based on a LangmuirHinshelwoodHougenWatson rate expression, using activity coefcients from the UNIFAC method, was used
to predict the experimental data of the heterogeneous liquid-phase reaction. The kinetic parameters were
determined based on the kinetic model. The model predicts the kinetic data very well and it will be useful
for design and optimization of integrated reactionseparation processes.
2014 Elsevier B.V. All rights reserved.

1. Introduction
In recent years, there has been a growing trend in the development of sustainable technology for production of environment
friendly gasoline and diesel fuels. Biodiesels or fatty acid alkyl
esters are an alternative fuel which is obtained from vegetable oils
or animal fats. Biodiesels have various technical advantages over
conventional gasoline or diesel fuels such as the reduction of emissions, enhanced lubricity and biodegradability, higher ash point
and lower toxicity [1]. They also show properties such as cetane
number, heat of combustion and viscosity that are similar to the
conventional diesels. However, they are inferior to conventional
diesel in terms of oxidation stability, nitrogen oxides emissions,
energy content and cold weather operability [2]. Employment of
additives is one of the approaches that can improve the quality
Corresponding author. Tel.: +351 225 081 671; fax: +351 225 081 674.
E-mail address: arodrig@fe.up.pt (A.E. Rodrigues).
http://dx.doi.org/10.1016/j.cej.2014.11.077
1385-8947/ 2014 Elsevier B.V. All rights reserved.

of diesel fuel [3]. A wide variety of additives are added to improve


fuel efciency and reduce harmful emissions. Nowadays, methyl
tert-butyl ether (MTBE) and ethyl tert-butyl ether (ETBE) are well
known oxygenated additives for gasoline. However, these ethers
are not suitable as they drastically reduce the cetane number in
diesel blends [4]. An alternate way to overcome this is to use acetal
as oxygenated additive [5,6]. Moreover, acetals nd their use in
raw materials for perfumes, agricultural chemicals and pharmaceutical applications.
Acetal can be produced from an acid-catalyzed reaction of 2 mol
of monohydric alcohol and 1 mol of aldehyde. The aldehydes can
be produced from their corresponding alcohols following a partial
oxidation or a dehydrogenation process. In other words, acetals can
be obtained from just renewable raw materials. The synthesis of
acetal was carried using homogeneous catalyst such as strong
liquid acids like H2SO4, HF and HCl [79]. But, their use suffers
from some drawbacks, mainly separation from the products due
to miscibility with the reaction mixture and corrosion of the equip-

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

Nomenclature
a
ap
Ap
C
Ca
Cb
C b0
Cp
C p0
C ps
De,j
Dj;m

liquid-phase activity
specic particle surface area (cm1)
total external area of particles (cm2)
concentration (mol cm3)
Carberry number
bulk concentration (mol cm3)
initial bulk concentration (mol cm3)
concentration inside the particle (mol cm3)
initial concentration inside the particle (mol cm3)
concentration on the particle surface (mol cm3)
effective diffusivity (cm2 min1)
molecular diffusivity in the mixture (cm2 min1)

D0j;i

binary diffusivity of reactant j in component i


(cm2 min1)
reaction activation energy (J mol1)
standard Gibbs free energy (J mol1)
standard Gibbs energy of formation (kJ mol1)

Ea
DG0
DG0f

DHs
DH0
DH0f
k0;c
kc
K 0;s
Keq
kL
Kc
Ks
KX
MRD
N
NE
NMi
S0
DS0
r
r A=B
r obs

enthalpy of adsorption of water (J mol1)


standard enthalpy (J mol1)
standard enthalpy of formation (kJ mol1)
pre-exponential constant (mol g1 min1)
kinetic constant based on LHHW kinetic model
(mol g1 min1)
constant in Eq. (29)
equilibrium reaction constant (dimensionless)
extraparticle mass transfer coefcient (cm min1)
equilibrium constants based on activity coefcients
(dimensionless)
adsorption constant (dimensionless)
equilibrium constants based on the molar fractions
(dimensionless)
mean relative deviation
total number of measurements
number of experiments performed
number of measurements of variable in the ith experiment
standard entropy of formation (J mol1 K1)
standard entropy of reaction (J mol1 K1)
radial position (cm)
initial ratio of reactants
observed reaction rate (mol g1 min1)

ment. The use of heterogeneous catalytic processes has overcome


most of the problems. Therefore, heterogeneous catalysts such as
ion-exchange resins, silica-based mesoporous materials and zeolites have become the alternate path for acetal production [5,10
12]. Among the heterogeneous catalysts, such as zeolites, depict
good performances; however, acid ion-exchange resins have been
shown very effective for this type of reaction and attaining higher
yields in short periods of time [5,12,13]. Several ion-exchange resins such as Amberlyst 15 [5,1316], Amberlyst 36 [15,17,18],
Amberlyst 46 [19,20], Amberlyst 70 [19,20], and Amberlyst 131
[19,21] were tested and studied by numerous researchers in the
applications of alkylation, etherication, esterication, transesterication and acetalization reactions.
Acetalization reactions are typically equilibrium limited reactions and the determination of thermodynamic equilibrium is
important. The probable mechanisms for heterogeneous kinetic
model of equilibrium limited reactions have been investigated by
many researchers. The pseudo-homogeneous model (PH) is widely
used in this kind of systems [22,23]. In the PH model, adsorption
and desorption of all compounds are neglected. On the other hand,
EleyRideal (ER) model can be applied when reaction between one
adsorbed reactant and one non-adsorbed reactant from the bulk

rp
R
Rc
R
Rp
t
T
V
V liq
Vp
x
X exp
X model
zik
~zik

259

particle radius (cm)


universal gas constant (J mol1 K1)
ratio of catalyst volume to catalyst external surface area
reaction rate (mol g1 min1)
rate of reaction relative to the local concentration
(mol g1 min1)
time coordinate (min)
temperature (K)
molar volume (mol cm3)
volume of liquid inside the reactor (cm3)
total volume of particles (cm3)
molar fractions
model predicted conversion
measured value of conversion
kth model predicted value of variable in experiment i
kth measured value of variable in experiment i

Greek letters
c
activity coefcients
eb
bulk porosity
ep
porosity of catalyst particle
h
set of model parameters to be estimated
l
viscosity (cp)
m
stoichiometric coefcient
q
dimensionless radial coordinate
qp
catalyst/particle density (g cm3)
s
tortuosity factor
/wp
WeiszPrater parameter
U
parameter in Eq. (26)
Subscripts
A
ethanol
B
butyraldehyde
C
DEB
D
water
exp
experimental
i
relative to component i
j
relative to component j
liq
liquid phase
m
mixture
model
predicted from model
p
particle
s
relative to the surface of the particle

liquid phase occur [24,25]. The LangmuirHinshelwoodHaugen


Watson (LHHW) model takes into account the adsorption of all
the compounds on catalyst surface. The LHHW model has already
been used to describe the reaction of acetaldehyde with several
linear chain alcohols [3,6,19,26].
There are some studies about the acetalization reaction
between ethanol and butyraldehyde in the literature. Arisketa
and co-workers [27,28] studied the kinetics of the synthesis of
1,1-diethoxybutane using Amberlyst 47. They have reported the
apparent value of rate constant and did not consider the non-ideality of the reaction mixture. The possibility of side reactions was
also reported in literature [27,28] but it has been found that the
amount of byproducts formed using Amberlyst 47 ion-exchange
resin catalyst was negligible.
In the present work, liquid-phase reaction of ethanol and butyraldehyde catalyzed by Amberlyst 47 for the synthesis of the acetal
1,1-diethoxybutane was studied in a batch reactor to obtain thermodynamic and kinetic data. Effects of temperature, catalyst loading and initial molar ratio of reactants on the reaction kinetics were
investigated. An activity based kinetic model was suggested to
describe the mechanism of heterogeneous catalytic liquid-phase
reaction.

260

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

2. Material and methods

3.1. Determination of equilibrium constant and reaction enthalpy

2.1. Chemicals and catalyst

It is very important to determine the equilibrium constant for


equilibrium limited reactions such as esterication, transesterication, acetalization, etc. To calculate the equilibrium constant,
experiments were undertaken to determine the equilibrium mole
fractions of ethanol, butyraldehyde, DEB and water. Experimental
runs were carried out in the temperature range 293.15323.15 K,
at 6.0 bar pressure, and initial molar ratio of ethanol to butyraldehyde of 2.1, and a catalyst loading of 1 wt.%. Experiments were performed at least three times to minimize the experimental errors.
All the experiments were performed for enough time so that the
reaction reached the equilibrium. Then, the equilibrium constant,
Keq, at different temperatures, for synthesis of DEB, was calculated
using the experimentally obtained data according to the following
formula:

Ethanol (99.5% purity) was obtained from Panreac. Butyraldehyde and 1,1-diethoxybutane (97% purity) were obtained from
Acros Organics. Acetone (99.98% purity) was obtained from Fisher
Chemical and used as the internal standard in gas chromatograph.
Deionized water was used for calibration purpose in gas chromatograph. A commercial strong acid ion-exchange resin named
Amberlyst 47 was used as the catalyst. The catalyst was provided
by Rohm and Haas (USA) in an opaque spherical bead-form.
Amberlyst 47 has a macroporous cross-linked polystyrene polymer
matrix with sulfonic groups according to the datasheet provided by
the manufacture. No information was available for the crosslinking
degree; however, it has a less to moderate swelling property in
methanol and water which means higher cross-linking degree
and rigid polymer structures. It has a surface area of 50 m2/g with
an average pore diameter of 240 .
2.2. Reactor setup
The experiments of reaction between ethanol and butyraldehyde
were carried out in a closed glass-jacketed1 dm3 autoclave reactor
(Bchi, Switzerland). The reactor was mechanically stirred and it is
equipped with pressure and temperature sensors, and with a
blow-off valve. The reactor was operated in a batch mode and its
temperature was controlled by thermostated silicon oil (Lauda, Germany) that ows through the jacket. The reactor was pressurized
with helium to maintain the reaction mixture in liquid phase over
the entire temperature range. The reactants were charged into the
reactor and heated to achieve the desired reaction temperature.
The dry catalyst was placed in a basket at the top of the stirrer shaft.
The basket falls down in the reaction solution as soon as the agitation
starts. In this way, the starting time of the reaction was accurately
dened. Samples were collected through one of the outlets of the
reactor in denite time interval and then analyzed.
2.3. Analytical method
All the samples were analyzed in a gas chromatograph (GC2010 Plus, Shimadzu) using a fused silica capillary column (CPWax 57 CB, 25 m  0.53 mm ID, df = 22.0 lm) to separate the compounds. The column was connected with a ame ionization detector (FID) and a thermal conductivity detector (TCD) to identify and
quantify the compounds present in the sample. Water was
detected in the TCD whereas the other compounds were identied
in FID. Helium N50 was used as the carrier gas. The injector and
detector temperature was maintained at 250 C. The column temperature program of gas chromatographic analysis was given as
follows: 7.5 min initial hold at 50 C; heating from 50 to 100 C
at a rate of 50 C/min and holding for 2 min. The reproducibility
of peaks was found to be good for all of the compounds in gas
chromatograph.
3. Thermodynamic analysis for the synthesis of 1,1diethoxybutane
The acetal, 1,1-diethoxybutane (DEB), is produced from the
homogeneous liquid phase reaction of ethanol and butyraldehyde
in a batch reactor described in the previous section. The reaction
is catalyzed by sulfonic acid ion exchange resin (Amberlyst 47)
according to the following stoichiometry.
H

2 Ethanol A Butyraldehyde B
DEB C Water D

K eq

aC aD xC xD cC cD


KXKc
a2A aB x2A xB c2A cB

where, a are respective activity of the compounds, x are molar fractions of the compounds and c are the activity coefcients of the
compounds at the equilibrium. KX and Kc are the equilibrium constants based on the molar fractions and activity coefcients, respectively. Due to strong non-ideality of the reaction mixture, activity
coefcients of compounds in the liquid phase were used to calculate
the equilibrium constant. The activity coefcients of compounds
were computed by the UNIFAC method [29,30]. The parameters
needed for UNIFAC method were taken from literature [31] and
they are presented in Appendix A. The average experimentally measured equilibrium compositions and the corresponding equilibrium
constant at each temperature are shown in Table 1.
The temperature dependence of equilibrium constant, Keq, is
given by,

ln K eq

D S0 DH 0 1

R
R T

where, DS0 and DH0 are the standard entropy and standard enthalpy
of the reaction, respectively, and R is the universal gas constant.
From the values of Keq at different temperatures, the standard
entropy and standard enthalpy of the reaction were determined
by plotting ln Keq vs. 1/T. The plot of ln Keq vs. 1/T is shown in
Fig. 1. The data was tted well by a straight line. The slope of this
straight line gives the value of the standard enthalpy of the reaction
while the standard entropy was calculated from the intercept of the
straight line. From the values of DS0 and DH0 , Eq. (3) can be
expressed as,

ln K eq

1036:80
 2:56
T

Table 1
Experimental equilibrium compositions and equilibrium constants.
Temperature
293.15 K

303.15 K

313.15 K

323.15 K

xA
xB
xC
xD

0.3772
0.1966
0.2131
0.2131

0.3856
0.2016
0.2064
0.2064

0.3960
0.2030
0.2005
0.2005

0.4076
0.2054
0.1935
0.1935

KX

1.6235

1.4212

1.2628

1.0972

cA
cB
cC
cD

1.1106
1.2768
1.3652
1.8863

1.1156
1.2799
1.3687
1.9372

1.1186
1.2831
1.3755
1.9809

1.1198
1.2853
1.3830
2.0228

Kc

1.6352

1.6646

1.6970

1.7357

Keq = KXKc

2.6547

2.3657

2.1430

1.9044

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

261

lyst 47 in a batch reactor are directly related to the stirrer speed


and the particle size of the catalyst, respectively.
4.1.1. External mass transfer resistance
In order to determine the inuence of external mass transfer
resistance, experiments at different stirring speeds were carried
out. These experiments were performed at the temperature of
323 K, initial molar ratio of ethanol to butyraldehyde of 2.1, catalyst loading of 1 wt.% and different stirring speeds to evaluate the
limits at which external diffusion limitations will not exist. The
effect of stirring speeds is shown in Fig. 2. It was found that there
was no limitation due to external resistance above 500 rpm, which
is also in agreement with the results reported in the literature [28].
Furthermore, existence of external mass transfer resistance was
investigated by Carberry number. The Carberry number Ca gives
the ratio between the observed reaction rate and maximum external mass transfer and is expressed by [33,34]:

Fig. 1. Linearization of the experimental equilibrium constants.

1

Therefore, the values of DH0 8619:96  287:22 J mol and


1
DS 21:28  0:93 J mol K1 were obtained. The error was
determined through least square method. The value of DH0 is
reported in the literature [27,28] as 31,098 J mol1. The experimentally determined value of DH0 in this present study is quite
less than the value reported in the literature. This may be due to
purely computation method employed in their work. On the other
hand, there was no data available for DS0 in any literatures. The
standard free energy change for the liquid phase reaction is related
to the standard enthalpy and entropy changes by,
0

DG0 DH0  T DS0

From the Eq. (5), the value of standard free energy change was
1
obtained as DG0 2275:33 J mol . Based on the experimental
values of DH0 , DS0 and DG0 , the standard formation properties of
DEB were calculated and presented in Table 2 [32]. The value of
standard enthalpy of formation of DEB given in literature as
540.46 kJ mol1 [27].

4. Kinetic studies
The experimental results of the reaction kinetics of the production of DEB from the reaction of ethanol and butyraldehyde catalyzed by the Amberlyst 47 resin are presented in this section.
The effect of various conditions, such as catalyst loading, initial
molar ratio between ethanol and butyraldehyde, and reaction temperature, on the conversion of butyraldehyde as a function of time
was studied. These studies were performed varying the condition
under evaluation and keeping constant the remaining conditions.

Ca

r obs qp C b  C ps

kL ap C b
Cb

where robs is the observed reaction rate, qp is the particle density,


kL is the external mass transfer coefcient, ap is the specic particle
surface area, C b is the bulk concentration of butyraldehyde and C ps
is the concentration of butyraldehyde on the particle surface. The
criteria for negligible external mass transfer resistance the value
of Ca should be less than 0.05 [34]. In this study the value of Ca
was found to be in an order of 103 for the experiment conducted
at 500 rpm. Therefore, all further experiments were performed at
500 rpm to ensure that the external mass transfer resistance does
not exist. Usually the rate of reaction in ion exchange catalyzed processes do not depend on the external mass transfer resistance
unless the viscosity of the reactant mixture is very high or the speed
of agitation is very low [14].
4.1.2. Internal mass transfer resistance
Different particle sizes of Amberlyst 47 resin are not commercially available, so it was not possible to investigate the internal
mass transfer limitation experimentally and thus, the Weisz
Prater criterion was employed. The WeiszPrater equation for
heterogeneous catalytic reaction can be expressed as in [35]:

/wp

r obs qP R2c
De;j C j

where /wp is the WeiszPrater parameter, Rc (=rp/3) is the ratio of


the catalyst volume to the catalyst external surface area where rp

4.1. Mass transfer limitation


To accomplish the true kinetic study for heterogeneous catalytic
system, it is necessary to evaluate both external and internal mass
transfer resistances. The external and internal mass transfer resistances of the reaction of ethanol with butyraldehyde over Amber-

Table 2
Standard formation properties at 298.15 K.

Component

DH0f kJ mol1

DG0f kJ mol1

S0 J mol

Ethanol [17]
Butyraldehyde [17]
DEB*
Water [17]

277.6
239.2
517.2
285.8

174.8
119.5
234.3
237.1

160.7
246.6
476.7
70.0

Calculated from this work.

1

K1

Fig. 2. Effect of stirring speeds on conversion of butyraldehyde (rA/B = 2.1,


catalyst = 1 wt.%, P = 6 bar, T = 323 K).

262

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

(=0.05 cm) is the particle radius [27], Cj is the concentration of limiting reactant in the mixture, and De,j is the effective diffusivity. The
value of qP is taken as 0.5508 g/cm3 [27]. The WeiszPrater criterion states that diffusion does not limit the reaction if the value of
/wp is less than 1. On the other hand, strong diffusion limitations
exist if /wp > 10. The expression for determination of effective diffusivity is given as,

De;j

ep Dj;m
s

where ep 0:5 is the porosity of catalyst particle [27], Dj;m is the


molecular diffusivity of limiting reactant in the mixture and s is
the tortuosity factor of catalyst. Estimation of the tortuosity was
performed using the correlation given by Wakao and Smith [36].
Dj;m was calculated using Perkin and Geankoplis [37] method for
multicomponent system

Dj;m l0:8
m

n
X

xi D0j;i l0:8
i

where xi is the mole fraction of the component i,


is the binary
diffusivity of reactant j in component i, and li and lm are the viscosities of component i and the mixture, respectively. The viscosity of
the liquid mixture, lm, was calculated by the Grunberg-Nisaan
method [38]. D0j;i was estimated by the following Scheibel correlation, which is the modied WilkeChang equation in order to eliminate the association factor [39]

"

1=3
iV j

3V i
Vj

2=3

#
10

where V is the molar volume of the respective compounds and T is


the temperature.
The WeiszPrater parameters, /wp , were estimated from some
of the experimental data and are presented in Table 3. The values
of /wp from the table indicate that the reaction of ethanol with
butyraldehyde using the ion exchange resin was limited by the
internal diffusion to some extent. Therefore, the effect of internal
diffusion was taken into account in further analysis of reaction
kinetics.
4.2. Kinetic model
It is evident from the values presented in Table 3 that the kinetics of the reaction is affected by the internal diffusion. Therefore, an
isothermally operated batch reactor model that considers diffusion
of the components inside the catalyst particle was used in this
study [16,40].
The mass balance in the bulk uid of batch reactor in the liquid
phase at constant temperature is given by,


dC b;j
Ap
@C p;j 

De;j
j A  D
dt
V liq
@r rrp

11

with

Ap



@C p;j
1 @
@C p;j
1  ep mj qp Rp
De;j r 2
2
r @r
@t
@r

3
Vp
rp

12

t 0 C b;j C b0;j ; C p;j C p0;j

Temperature (K)

/wp

293.15
303.15
313.15
323.15

1.52
1.76
2.18
3.03

14

Considering negligible external mass transfer resistance, the


boundary conditions are,

r rp

@C p;j
0
@t

15


C p;j C p;j rrp

16

Introducing the dimensionless space variable q r=rp , the


model equations can be expressed as,


dC b;j
3 1  eb
@C p;j 
 2
De;j
rp eb
dt
@ q q1

17



@C p;j De;j 1 @
@C
1  ep
2 2
mq R
q2 p;j
@t
rp q @ q
@q
ep j p p

18

where eb is the bulk porosity. The boundary conditions for above


model equations then become:

q0

@C p;j
0
@t

19


q 1 C p;j C p;j q1

20

It should be noted that the amount of catalyst was implemented


in the kinetic model through the calculation of the total volume of
particles, Vp, using the value of catalyst density; then the total
interfacial area uid/particle Ap was calculated from Eq. (12).
In this study, the LangmuirHinshelwoodHaugenWatson
model equation was used. It was considered to be the most appropriate model for the reaction between ethanol and butyraldehyde
catalyzed by Amberlyst 47, following the previous works done
with diethylacetal and dimethylacetal synthesis [16,26,40]. Therefore the present model was developed on the basis of LHHW mechanism. The mechanism is based on adsorption of the reactant
species (ethanol and butyraldehyde), the reaction between
adsorbed reactants on the catalyst surface, and desorption of the
reaction products (DEB and water). The surface reaction involves
the following steps:
Surface reaction between adsorbed species of ethanol (A) and
butyraldehyde (B) to give absorbed hemiacetal, I1  S:

A  S B  S
I1  S S
Table 3
WeiszPrater parameter for determination of internal
diffusion
(rA/B = 2.1,
catalyst = 1 wt.%,
stirring
speed = 500 rpm, P = 6 bar).

13

where mj is the stoichiometric coefcient of the component j and Rp


is the rate of reaction relative to the local concentration.
Initial conditions:

r0
D0j;i

8:2  108 T

ep

i1
ij

D0j;i

where C b;j is the bulk concentration of component j, C p;j is the concentration of component j inside the particle pore, V liq is the volume
of liquid inside the reactor, r p is the particle radius, Ap is the total
external area of particles, V p is the total volume of particles, r is
the radial position and t is the time coordinate.
Similarly, mass balance inside the catalyst particle at constant
temperature is given by,

21

Surface reaction to obtain absorbed water, D  S:

I1  S S
I 2  S D  S

22

Surface reaction to obtain adsorbed DEB, C  S:

I2  S A  S
C  S S

23

The surface reaction is assumed to be the controlling step and


also multicomponent Langmuir adsorption isotherms are assumed
for describing the adsorption behavior of the compounds of the

263

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

reaction mixture in the surface of the resin. The model is based on


activities (ai) of the respective compounds. Taking into account the
above considerations, the following rate expression, considering
nonideality of the reaction mixture, is obtained [26]:

R kc

aA aB  aaACKaeqD
1 K s;A aA K s;B aB K s;C aC K s;D aD K I1 aA aB K I2 aC =aA 2
24

where R is the reaction rate, kc is the kinetic constant based on the


kinetic model and K s are the adsorption constant for the respective
components. A detailed derivation of the rate expression is
described in Appendix B. In order to reduce the number of optimization parameters, only those compounds which had strongest
adsorption were taken into consideration. Water has the strongest
adsorption strength on the resin surface as the acidic property of
Amberlyst 47. Thus, the adsorption of the other compounds was
neglected. The simplied rate expression which was used to
describe the experimental data is:

R kc

aA aB  aaACKaeqD
1 K s;D aD 2

25

According to the above kinetic model, there were two parameters to be estimated, the kinetic constant kc and the adsorption
constant of water K s;D ; at each temperature.
4.2.1. Parameter estimation from experimental data
The model equations were solved numerically using the software gPROMS (general PROcess Modeling System), version 3.3.1
which is commercial package from Process System Enterprise.
The particle radial domain in batch reactor model was discretized
using the second-order orthogonal collocation in the nite
element method (OCFEM). The system of ordinary differential
equation was integrated over time using DASOLV integrator
implemented in gPROMS. For radial discretization, 10 nite
elements with two collocation points were used in each element.
For all simulations, tolerance equal to 105 was xed. The
unknown parameters kc ; K s;D proposed in the kinetic model were
estimated using Parameter Estimation tool in gPROMS providing the best t of measured and predicted conversion data using
the maximum likelihood method. The objective function associated with the parameter estimation is described by the following
equation:

( NM "
#)
2
NE Xi
X
N
1
~zik  zik
U ln2p min
lnr2ik
2
2 h
r2ik
i1 k1

butyraldehyde (2.14.1), and also at various catalyst loading


(0.51.5%). For the proposed kinetic model, the unknown parameters kc ; K s;D were estimated from the conversion data along time
for the reaction between ethanol and butyraldehyde using Amberlyst 47 resin as catalyst. The optimized values of these parameters,
at each temperature, are presented in Table 4.
The temperature dependence of kinetic and adsorption constants were described by the Arrhenius and Vant Hoff equations,
respectively:



Ea
kc k0;c 
RT

28



DH s
K s;D K 0;s 
RT

29

where Ea is the reaction activation energy, k0;c is the pre-exponential constant, R is the universal gas constant, K 0;s is a constant,
DHs is the enthalpy of adsorption of water and T is the temperature.
The plot of kc and K s;D tted in Arrhenius and Vant Hoff equation, respectively, are shown in Fig. 3. The data were tted well
by the straight lines. The slopes of these lines give the values of
the activation energy Ea and enthalpy of adsorption DHs . The
corresponding values were obtained as Ea 37:92  2:58 kJ=mol
and DHs 24:28  1:8 kJ=mol, respectively. A value of
Ea = 35.50 kJ/mol was obtained by Arisketa and co-workers
[27,28]. This may be due to consideration of a non-ideal mixture
and a simple kinetic model to describe the acetalization reaction.
A summary of the proposed kinetic model and their parameters
is presented in Table 5. In order to validate the existence of internal
diffusion resistance; by simulation it is possible to observe the concentration gradient inside the catalyst particle. Fig. 4 shows the
internal concentration prole of butyraldehyde at different temperatures after 10 min. The presence of concentration gradient
between the surface and the center of the catalyst indicates the
presence of internal mass transfer resistance.

Table 4
Estimated model parameters at different temperatures.
Temperature (K)
293.15
1

kc mol g
K s;D

min

1

303.15

313.15

323.15

0.06196

0.08603

0.15584

0.25050

2.9385

2.2377

1.6991

1.1566

26

where N is the total number of measurements taken during all the


experiments, h is the set of model parameters to be estimated, NE is
the number of experiments performed, NMi is the number of measurements of variable in the ith experiment, r2ik is the variance of
the kth measurement of variable in experiment i, ~zik is the kth measured value of variable in experiment i and zik kth model predicted
value of variable in experiment i. The quality of the model t was
tested through the mean relative deviation (MRD) between the data
of model predicted conversion X model and the experimental data
X exp using the following equation.

MRD%




1 X X model  X exp 

  100
N
X exp

27

4.3. Modeling and discussion of results


Experiments were performed at different temperatures
(293.15 K323.15 K), at various initial mole ratio of ethanol to

Fig. 3. Arrhenius and Vant Hoff plots for kc and Ks,D, respectively, at different
temperatures.

264

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

Table 5
Kinetic law and parameters from proposed kinetic model.
Kinetic law

R kc faA aB  aC aD =aA K eq g=1 K s;D aD 2 

Equilibrium constant
(dimensionless)
Kinetic constant
(mol g1 min1)
Adsorption constant of water
(dimensionless)

K eq 7:73  102 exp1036:80=TK


kc 3:21  105 exp4561:06=TK
K s;D 1:454  104 exp2920:68=TK

Fig. 6. Effect of catalyst loading on conversion of butyraldehyde (rA/B = 2.1, stirring


speed = 500 rpm, P = 6 bar, T = 303 K).

Fig. 4. Internal concentration prole of butyraldehyde after 10 min.

Fig. 7. Effect of initial molar ratio of ethanol and butyraldehyde on conversion of


butyraldehyde (catalyst = 1 wt.%, stirring speed = 500 rpm, P = 6 bar, T = 303 K).

Fig. 5. Effect of reaction temperature on conversion of butyraldehyde (rA/B = 2.1,


catalyst = 1 wt.%, stirring speed = 500 rpm, P = 6 bar).

4.3.1. Effect of reaction temperature


The effect of temperature on the conversion of butyraldehyde is
shown in Fig. 5. It is evident from the gure that the rate of reaction increases with temperature while the equilibrium conversion
of butyraldehyde decreases due to the exothermic nature of this
reaction. It is also observed from the gure that the experiments
are well predicted by the model.
Fig. 8. Parity plot comparing experimental and calculated conversions.

4.3.2. Effect of catalyst loading


Fig. 6 shows the effect of the catalyst loading on the conversion
of butyraldehyde at 303.15 K, where it can be seen that the reaction rate increases with increase in catalyst loading. This is due
to the fact that the increase in catalyst loading results in higher
number of active catalytic sites available for reaction. Once again,
the model was able to predict the experimental conversion data
for various catalyst loadings.

4.3.3. Effect of initial molar ratio of reactants


As mentioned above, the effect of initial mole ratio of ethanol to
butyraldehyde was studied by varying it from 2.1 to 4.1. The
obtained results are presented in Fig. 7. It is observed in this gure
that, as expected, the equilibrium conversion increases for higher
values of initial mole ratio. This behavior is well described by
the developed model. However, a small overestimation of model

265

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

where k1 and k1 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as
q
a2
K 0s;A kk11 a2ASa2 with K s;A K 0s;A aaAASaS :

predicted conversion over the experimental conversion was found


for the initial molar ratio of 4.1.
The proposed model describes very well the experimental data
at various temperature, catalyst loading and initial molar ratio of
the reactants up to the equilibrium conversion and also the equilibrium stage (Figs. 57). A comparison of the experimental data
of conversion of butyraldehyde and those predicted from the
model is shown in Fig. 8. The value of MRD between experimental
and model-predicted conversions of 7.7% was obtained.

where k2 and k2 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as

5. Conclusion

K s;B kk22 aaBBSaS :

The synthesis of 1,1-diethoxybutane in a liquid phase reaction


of ethanol and butyraldehyde catalyzed by Amberlyst 47 was studied in a batch reactor. The thermodynamic equilibrium constant
was determined in the temperature range of 293.15 K323.15 K
and can be expressed as K eq 7:73  102 exp1036:80=TK.
The standard properties of the reaction at 298.15 K are
DH0 8619:96
J mol1,
DS0 21:28
J mol1 K1
and
0
1
DG 2275:33 J mol . Due to the strong non-ideality of the
liquid reaction mixture, the kinetic model was formulated in terms
of activities. A two-parameter kinetic model based on a Langmuir
HinshelwoodHaugenWatson rate expression was proposed to
describe the experimental kinetic results. The effects of different
parameters such as temperature, catalyst loading, and ethanol to
butyraldehyde initial mole ratio were studied. The model parameters such as experimental rate constant and adsorption coefcient
of water were determined. The simulation results were in good
agreement with the experimental results and this model can be
used to predict the batch reactor performance for the synthesis
of 1,1-diethoxybutane. These data will be useful for the implementation of integrated reactionseparation processes such as xed
bed reactors and simulated moving bed reactors to improve the
conversion to 1,1-diethoxybutane.

A S

Step 2 adsorption of butyraldehyde:


k2

BS BS

B2

k2

Step 3 surface reaction between the adsorbed species of ethanol and aldehyde to give adsorbed hemi-acetal:
k3

A  S B  S I1  S S

B3

k3

where k3 and k3 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as
aI

S aS

K 3 kk33 aAS1 aBS :


Step 4 surface reaction to obtain adsorbed water:
k4

D  S : I1  S S I2  S D  S
k4

where k4 and k4 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as
K 4 kk44

aI

a
2 S DS

aI

a
1 S S

Step 5 surface reaction of formation of adsorbed DEB:


k5

C  S : I2  S A  S C  S S
k5

aS
K 5 kk55 aaI CS
:
S aAS
2

Step 6 desorption of DEB:


k6

B6

k6

Financial support for this work was provided by project PEst-C/


EQB/LA0020/2011, nanced by FEDER through COMPETE Programa Operacional Factores de Competitividade and by FCT Fundaopara a Cincia e a Tecnologia, for which the authors are
thankful. This work was also co-nanced by QREN, ON2 and FEDER
(Project NORTE-07-0124-FEDER-0000007 Multifunctional Reactors/Process Intensication). C.S.M. Pereira acknowledges Fundo
Social Europeu (European Social Fund) and Programa Operacional
Potencial Humano (Human Potential Operational Programme)
(FCT Investigator-IF/01486/2013).This project has been also
funded with support from the European Commission. This communication reects the views only of the authors, and the Commission
cannot be held responsible for any use which may be made of the
information contained therein.

Appendix A.
Tables A.1 and A.2 present the UNIFAC method parameters used
in this work to calculate the activity coefcients.
Appendix B.

B5

where k5 and k5 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as

CS CS
Acknowledgments

B4

where k6 and k6 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as
K s;C kk66 aaCCSaS :
Step 7 desorption of water:
k7

DS DS

B7

k7

where k7 and k7 are the forward and backward rate constant,
respectively. The adsorption coefcient can be written as
K s;D kk77 aaDDSaS :
Assuming the surface reaction (i.e., Step 4) is the rate
determining step, the rate expression for LangmuirHinshelwoodHaugenWatson (LHHW) kinetic model is:


 

k4
hI2 S hDS
R k4 hI1 S hS  k4 hI2 S hDS k4 hI1 S hS 
k4

 

1
k4 hI1 S hS 
hI S hDS
K4 2

B8

where hI1 S ; hI2 S and hDS are the fractions of catalyst sites occupied
by I1  S, I2  S and D  S, respectively, hS is the fraction of vacant sites
and K 4 is the adsorption coefcient for step 4.
The total activity composed of vacant and the adsorbed species
on the catalyst surface can be expressed as:

a0 aS aAS aBS aCS aDS aI1 S aI2 S


The kinetic model described in this work based on the activity
of the respective species and has the following steps:
Step 1 adsorption of ethanol:
k1

2A 2S 2A  S
k1

B1

aS K s;A aA aS K s;B aB aS K s;C aC aS K s;D aD aS


K

a a

s;C C S
K s;A K s;B K 3 aA aB aS K s;A
K 5 aA
h
i
K a
aS 1 K s;A aA K s;B aB K s;C aC K s;D aD K s;A K s;B K 3 aA aB K s;As;CK 5CaA

B9

266

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267

Table A.1
Relative molecular volume and surface parameters of the pure species.
Molecule (i)

Group identication
No. of main group

No. of secondary group

v i
k

Rk

Qk

Ethanol

CH3
CH2
OH

1
1
5

1
2
15

1
1
1

0.9011
0.6744
1.0000

0.848
0.540
1.200

Butyraldehyde

CH3
CH2
CHO

1
1
10

1
2
21

1
2
1

0.9011
0.6744
0.9980

0.848
0.540
0.948

DEB

CH3
CH2
CH
CH2O

1
1
1
13

1
2
3
26

3
2
1
2

0.9011
0.6744
0.4469
0.9183

0.848
0.540
0.228
0.780

Water

H2O

17

0.9200

1.400

where K 0I1

Table A.2
Interaction parameters for UNIFAC method.
am,n

Taking K I1

am,n

1
5
7
10
13

10

13

0
156.4
300
505.7
83.36

986.5
0
229.1
529
237.7

1318
353.5
0
480.8
-314.7

677
203.6
116
0
7.838

251.5
28.06
540.5
304.1
0

Taking hj as the fraction of catalyst sites occupied by a particular


species, can be expressed as:

hj

aI1 S aAS

a2A aB a2S

ajS
a0

K 0I1

K s;A

a2AS
a2A a2S



aI1 S aS
aAS aBS

K 2s;A K s;B K 3

B15

Similarly, adding Eqs. (B1), (B5) and (B6), we have:


k0I

I2  S 2A AS C
where

K 0I2

aCS aS

B16

a2AS

K 5 K 2s;A
aI S aAS a2 a2
aAS aC

2 aCS A S
2
K s;C
aI2 S aA
aC aS

Taking K I2

B10

K s;A
K s;C

K 5 K s;A
K 0I2

B17

Using the values of K I1 , K I2 and K eq , Eq. (B12) can be rearranged


as:


 

aI S aS
1 aI2 S aDS
R k4 1

K 4 a0 a0
a0 a0

 

K 3 aAS aBS aS
1
aCS aS aDS
k4

K 4 K 5 a0 aAS a0
a0 aS a0

 

K 3 K s;A K s;B aA aB a2S
1 K s;C K s;D aC aD a2S
k4

2
2
K4
K s;A K 5 aA a0
a0

 

1 K s;C K s;D aC aD aS 2
k4 K s;A K s;B K 3 aA aB 
a0
K4
K s;A K 5 aA
 
1 K s;C K s;D aC aD
K s;A K s;B K 3 aA aB 
K4
K s;A K 5 aA
k4 
2
K s;C aC
1 K s;A aA K s;B aB K s;C aC K s;D aD K s;A K s;B K 3 aA aB
K s;A K 5 aA
!
K s;C K s;D
1
aC aD
aA aB 
K 2s;A K s;B K 3 K 4 K 5 aA
k4 K s;A K s;B K 3 
2
K s;C aC
1 K s;A aA K s;B aB K s;C aC K s;D aD K s;A K s;B K 3 aA aB
K s;A K 5 aA

B11

Now the overall reaction can be expressed by:


keq

2A B C D

B12

where the overall equilibrium constant K eq is,

aCS
aDS
2
aC aD aS K s;C aS K s;D K s;A K s;B aCS aDS
2 2

aS
K s;C K s;D a2AS aBS
aA aB
aAS
aBS
aK
a2S K 02
s;A S s;B

aBS
aB aS

K s;A K s;B K 3

From the Eq. (B8) using the corresponding values of hj we have,

K eq



K 2s;A K s;B
K3K4K5
K s;C K s;D

B13

Adding Eqs. (B1)(B3), we get:


k0I

2A B 2S I1  S A  S

B14

R kc h

aA aB  KaeqC aaDA
1 K s;A aA K s;B aB K s;C aC K s;D aD K I1 aA aB K I2 aaCA

i2
B18

where kc k4 K s;A K s;B K 3


References
[1] F. Ma, M.A. Hanna, Biodiesel production: a review, Bioresour. Technol. 70
(1999) 115.
[2] B.R. Moser, S.Z. Erhan, Branched chain derivatives of alkyl oleates: tribological,
rheological, oxidation, and low temperature properties, Fuel 87 (2008) 2253
2257.
[3] G. Knothe, J.H. Van Gerpen, J. Krahl, The Biodiesel Handbook, AOCS press,
Champaign, IL, 2005.
[4] T. Li, M. Suzuki, H. Ogawa, Effects of ethyl tert-butyl ether addition to diesel
fuel on characteristics of combustion and exhaust emissions of diesel engines,
Fuel 88 (2009) 20172024.
[5] M.a.R. Capeletti, L. Balzano, G. de la Puente, M. Laborde, U. Sedran, Synthesis of
acetal (1,1-diethoxyethane) from ethanol and acetaldehyde over acidic
catalysts, Appl. Catal. A Gen. 198 (2000) L1L4.
[6] D. Naegeli, L. Dodge, Combustion Characterization of Methylal in Reciprocating
Engines, Final Report No. SwRI (1992) 03130.
[7] M. El-Chahawi, M. Kaufhold, Process for Preparing Acetaldehyde Diethyl
Acetal, Google Patents, 1996.
[8] J. Lilja, D.Y. Murzin, T. Salmi, J. Aumo, P. Mki-Arvela, M. Sundell, Esterication
of different acids over heterogeneous and homogeneous catalysts and
correlation with the Taft equation, J. Mol. Catal. A Chem. 182 (2002) 555563.
[9] A.K. Kolah, N.S. Asthana, D.T. Vu, C.T. Lira, D.J. Miller, Reaction kinetics of the
catalytic esterication of citric acid with ethanol, Ind. Eng. Chem. Res. 46
(2007) 31803187.
[10] J. Andrade, D. Arntz, M. Kraft, G. Prescher, Method for Preparation of Acetals,
Google Patents, 1986.
[11] V.M. Silva, A.E. Rodrigues, Novel process for diethylacetal synthesis, AIChE J. 51
(2005) 27522768.
[12] R.P. Faria, C.S. Pereira, V.M. Silva, J.M. Loureiro, A.E. Rodrigues, Glycerol
valorization as biofuel: thermodynamic and kinetic study of the acetalization
of glycerol with acetaldehyde, Ind. Eng. Chem. Res. 52 (2013) 15381547.
[13] S.M. Mahajani, Reactions of glyoxylic acid with aliphatic alcohols using
cationic exchange resins as catalysts, React. Funct. Polym. 43 (2000) 253268.

M. Rahaman et al. / Chemical Engineering Journal 264 (2015) 258267


[14] A. Chakrabarti, M. Sharma, Cationic ion exchange resins as catalyst, React.
Polym. 20 (1993) 145.
[15] E.. Akbay, M.R. Altokka, Kinetics of esterication of acetic acid with n-amyl
alcohol in the presence of Amberlyst-36, Appl. Catal. A 396 (2011) 1419.
[16] G.K. Gandi, V.M. Silva, A.E. Rodrigues, Process development for dimethylacetal
synthesis: thermodynamics and reaction kinetics, Ind. Eng. Chem. Res. 44
(2005) 72877297.
[17] B. Schmid, M. Dker, J. Gmehling, Esterication of ethylene glycol with acetic
acid catalyzed by Amberlyst 36, Ind. Eng. Chem. Res. 47 (2008) 698703.
[18] J. Deutsch, A. Martin, H. Lieske, Investigations on heterogeneously catalysed
condensations of glycerol to cyclic acetals, J. Catal. 245 (2007) 428435.
[19] T. Komon, P. Niewiadomski, P. Oracz, M.E. Jamrz, Esterication of acrylic acid
with 2-ethylhexan-1-ol: thermodynamic and kinetic study, Appl. Catal. A 451
(2013) 127136.
[20] C. Casas, R. Bringu, E. Ramrez, M. Iborra, J. Tejero, Liquid-phase dehydration
of 1-octanol, 1-hexanol and 1-pentanol to linear symmetrical ethers over ion
exchange resins, Appl. Catal. A 396 (2011) 129139.
[21] S. Karakus, E. Sert, A.D. Buluklu, F.S. Atalay, Liquid phase esterication of
acrylic acid with isobutyl alcohol catalyzed by different cation exchange
resins, Ind. Eng. Chem. Res. 53 (2014) 41924198.
[22] Y. Seo, W.H. Hong, Kinetics of esterication of lactic acid with methanol in the
presence of cation exchange resin using a pseudo-homogeneous model, J.
Chem. Eng. Jpn. 33 (2000) 128133.
[23] A. Izci, F. Bodur, Liquid-phase esterication of acetic acid with isobutanol
catalyzed by ion-exchange resins, React. Funct. Polym. 67 (2007) 14581464.
[24] Y. Liu, E. Lotero, J.G. Goodwin Jr, A comparison of the esterication of acetic
acid with methanol using heterogeneous versus homogeneous acid catalysis, J.
Catal. 242 (2006) 278286.
[25] S.R. Kirumakki, N. Nagaraju, K.V. Chary, Esterication of alcohols with acetic
acid over zeolites Hb, HY and HZSM5, Appl. Catal. A 299 (2006) 185192.
[26] V.M.T.M. Silva, A.E. Rodrigues, Synthesis of diethylacetal: thermodynamic and
kinetic studies, Chem. Eng. Sci. 56 (2001) 12551263.
[27] M.I.A. Arisketa, Innovative reaction systems for acetal (1,1 diethoxy butane)
production from renewable sources (Ph.d. thesis), University of the Basque
Country, 2010.

267

[28] I. Agirre, V.L. Barrio, B. Gemez, J.F. Cambra, P.L. Arias, Bioenergy II: the
development of a reactive distillation process for the production of 1, 1
diethoxy butane from bioalcohol: kinetic study and simulation model, Int. J.
Chem. Reactor Eng. 8 (2010).
[29] A. Fredenslund, R.L. Jones, J.M. Prausnitz, Group-contribution estimation of
activity coefcients in nonideal liquid mixtures, AIChE J. 21 (1975) 10861099.
[30] A. Fredenslund, J. Gmehling, P. Rasmussen, Vapor-Liquid Equilibria Using
UNIFAC: A Group Contribution Method, Elsevier Scientic Pub. Co.,
Amsterdam and New York, 1977.
[31] B.E. Poling, J.M. Prausnitz, O.C. John Paul, R.C. Reid, The Properties of Gases and
Liquids, McGraw-Hill, New York, 2001.
[32] D.R. Lide, CRC Handbook of Chemistry and Physics, CRC press, 2004.
[33] J.J. Carberry, Chemical and Catalytic Reaction Engineering, McGraw-Hill, New
York, 1976.
[34] F. Dekker, A. Bliek, F. Kapteijn, J. Moulijn, Analysis of mass and heat transfer in
transient experiments over heterogeneous catalysts, Chem. Eng. Sci. 50 (1995)
35733580.
[35] P. Weisz, C. Prater, Interpretation of measurements in experimental catalysis,
Adv. Catal. 6 (1954) 143.
[36] N. Wakao, J. Smith, Diffusion in catalyst pellets, Chem. Eng. Sci. 17 (1962) 825
834.
[37] L. Perkins, C. Geankoplis, Molecular diffusion in a ternary liquid system with
the diffusing component dilute, Chem. Eng. Sci. 24 (1969) 10351042.
[38] L. Grunberg, A.H. Nissan, Mixture law for viscosity, Nature 164 (1949) 799
800.
[39] E.G. Scheibel, Correspondence. Liquid diffusivities. Viscosity of gases, Ind. Eng.
Chem. 46 (1954) 20072008.
[40] N.S. Graa, L.s.S. Pais, V.M. Silva, A.E. Rodrigues, Oxygenated biofuels from
butanol for diesel blends: synthesis of the acetal 1,1-dibutoxyethane catalyzed
by amberlyst-15 ion-exchange resin, Ind. Eng. Chem. Res. 49 (2010) 6763
6771.

S-ar putea să vă placă și