Вы находитесь на странице: 1из 12

Chemical Engineering Science 62 (2007) 907 – 918

www.elsevier.com/locate/ces

Acetaldehyde dimethylacetal synthesis with Smopex 101 fibres as


catalyst/adsorbent
Ganesh K. Gandi, Viviana M.T.M. Silva, Alírio 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

Received 6 February 2006; received in revised form 7 October 2006; accepted 17 October 2006
Available online 28 October 2006

Abstract
The synthesis of dimethylacetal using acetaldehyde and methanol as raw material in the presence of Smopex 101 fibres as catalyst and
adsorbent in batch reactor and in a fixed bed adsorptive reactor, respectively, was studied. In the batch reactor the determination of thermodynamic
and reaction kinetics data for acetalization reaction was presented. A kinetic model based on a pseudo-homogeneous rate expression using
activities was proposed to describe the experimental kinetic results. The dynamic binary adsorption experiments were carried out in the absence
of reaction at 293.15 K in a laboratory scale column. The experimental results of the adsorption of binary non-reactive mixtures are reported
and used to obtain multi-component adsorption equilibrium isotherms of Langmuir type. The mathematical model was proposed to describe the
adsorptive reactor dynamic behaviour. The experimental results obtained for the reaction and regeneration experiments were compared with the
model proposed. Model equations were solved by orthogonal collocation on finite elements (OCFE) implemented by the PDECOL package,
using the measured model parameters and was validated for both reaction and regeneration steps.
䉷 2006 Elsevier Ltd. All rights reserved.

Keywords: Acetaldehyde dimethylacetal; Adsorption; Batch catalytic reactor; Fixed bed reactor; Kinetics; Smopex 101 fibres; Thermodynamics

1. Introduction et al., 2002), Zeolite-T membrane (Tanaka et al., 2002), ZSM-


5 (Ma et al., 1996; Yadav and Kishnan, 1998), HY Zeolite
The reaction between an alcohol and aldehydes to form an (Ma et al., 1996), Zeolite beta (Silva-Machado et al., 2000),
acetal and water is of considerable industrial interest. Acetals acid treated clays (Yadav and Kishnan, 1998), heteropoly-
are important class of chemicals having applications in a va- acids (Schwegler and van Bekkum, 1991; Lacaze-Dufaure and
riety of areas in the chemical industry such as perfumes, fla- Moulougui, 2000; Dupont et al., 1995). Ion exchange resins,
vors, pharmaceuticals, plasticizer, solvents and intermediates however, are the most commonly used solid catalysts and they
(Aizawa et al., 1994; Bauer et al., 1990; Gupta, 1987). have been proved to be effective in liquid phase acetalization
Most of the industrial reaction pertaining to the chemical, (Roy and Bhatia, 1987; Chopade and Sharma, 1996, 1997;
petrochemical and pharmaceutical sectors and almost all bio- Yadav and Pujari, 1999; Fedriani et al., 2000; Silva and
logical reactions are catalytic in nature. Several reviews have Rodrigues, 2001; Gandi et al., 2005). Typical resin catalysts are
recently been published on the use of solid acids as heteroge- sulphonic acids fixed to polymer carriers, such as polystyrene
neous catalysts for the preparation of specialty and fine chem- cross-linked with divinylbenzene (DVB). Solid ion-exchange
icals. Heterogeneous catalysts have been developed in order resins as catalyst have several advantages: the catalyst can
to avoid separation problems; a catalyst-free product can be be removed from the reaction mixture, continuous operation
easily obtained by filtration (Kekre and Gopala Rao, 1969). in column reactors is enabled, the product purity is high
There exist many solid catalysts, such as iodine (Ramalinga (Chopade and Sharma, 1996).
Heterogeneous catalysis compared with homogeneous catal-
∗ Corresponding author. Tel.: +351 22 5081671; fax: +351 22 5081674. ysis arises from the complex nature of the heterogeneous
E-mail address: arodrig@fe.up.pt (A.E. Rodrigues). process, which includes adsorption, desorption as well as
0009-2509/$ - see front matter 䉷 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.10.014
908 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

several surface reaction steps (Doraiswamy and Sharma, 1987). Table 1


However, application of adsorption, desorption and surface re- Physical properties of Smopex 101
action relationships to heterogeneous catalysed reactions is of Property Smopex 101
particular interest, since it is possible to obtain information on
Physical state Fibres
the reactivity of molecules in a specific reaction. Particularly
Colour Yellow
analysis of kinetic and thermodynamic results can improve the pH Acidic in water
knowledge of the reaction mechanisms on solid catalysts as Solubility in water Insoluble
well as the nature of the active sites of the solid catalysts. Mean fibre diameter 12.1 m
The specific advantage of fibre catalyst is due to negligible Average length 0.25 mm
Bulk density (packed bed) 410 ± 10 kg/m3
diffusional paths, which provide high catalyst efficiency and ex-
Fibre capacity 2.9 mmol/g
cellent separation abilities. The novel polymer-supported fibre
catalyst Smopex 101 (Smoptech Ltd., Finland) can be used in
various industrially relevant heterogeneous reactions, such as
esterification (Maki-Arvela et al., 1999; Lilja et al., 2002a,b),
etherification (Kariner et al., 2000; Kariner and Krause, 2001;
Pääkkönen and Krause, 2003), aldolization and hydrogenation
(Aumo et al., 2002). The polyethylene-based fibre catalyst can
be modified by grafting different functional groups, such as
pyridine, carboxylic acid, and sulphonic acid or a combination
of them. In this way homogeneous catalysts can be replaced by
fibre catalysts.
The conversion in acetalization reaction is limited by the
chemical equilibrium; therefore, the technology for the indus-
trial preparation of acetals involves the reaction and then the
separation of products from the reaction mixture at equilibrium
containing both products and reactants. The disadvantage of
this two-step procedure for operating equilibrium reactions is
in its economy, because of the high energy costs and the use of
different reactors and separators.
The objective of the present work is to demonstrate the ap-
plication of polymer supported sulphonic acid catalyst Smopex
101 (Smoptech Ltd., Finland) for the process development of
the acetalization reaction of methanol and acetaldehyde. The
study involves the measurement of acetalization kinetics and
thermodynamics and, thereafter, the analysis of the dynamic be-
haviour of acetalization reaction in a fixed bed chromatographic
reactor packed with Smopex 101 fibres for acetalization.

2. Chemicals used and catalyst

Acetaldehyde and acetaldehyde from Sigma-Aldrich were


used in this work in order to produce acetal (DME). The authen-
tic sample of acetal (DME) from Sigma-Aldrich was also used
for calibration of the Gas Chromatograph. Smopex 101 fibre
catalyst was obtained from Smoptech Ltd. (Johnson Matthey
Company, Finland). Smopex 101 is a poly(ethylene) based fi-
bre that is grafted with styrene, which is sulphonated with
chloro-sulphonic acid. These Smopex 101 fibres are used in
this present work as catalyst and as well as an adsorbent. Fig. 1. SEM Image of the Smopex 101 fibres: (a) fibre physical appearance;
The physical properties of Smopex 101 fibres are given in the (b) fibre cross section.
Table 1 as reported by Smoptech, Finland. The fibrous polymer-
supported sulphonic acid catalyst (Smopex 101) was investi-
gated with SEM (scanning electron microscopy); typical SEM
images are shown in Figs. 1(a and b). Fig. 1(a) shows the physi- the reaction occurs at the catalyst surface and, due to the ab-
cal appearance of Smopex 101 fibres, revealing averages diam- sence of internal diffusional resistances in the Smopex fibres,
eter and length of 12.1 and 250 m, respectively. The evidence the efficiency of this catalyst should be naturally higher than
that the fibres are non-porous is shown in Fig. 1(b). Therefore, other commercial catalysts, as Amberlyst 15.
G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918 909

3. Batch reactor Table 2


Effect of the temperature at the thermodynamic equilibrium
The thermodynamics and kinetic experiments for acetal T(K) 283.15 293.15 303.15 313.15
(DME) production were carried out in a batch reactor, at con-
xAe 0.308 0.313 0.316 0.335
trolled temperature and pressure (Silva and Rodrigues, 2001;
x Be 0.129 0.131 0.152 0.161
Gandi et al., 2006). x Ce 0.281 0.278 0.266 0.252
x De 0.281 0.278 0.266 0.252
3.1. Reaction thermodynamics Kx 6.428 5.998 4.662 3.543

A 0.859 0.873 0.863 0.873


The equilibrium constants were evaluated at several temper- B 0.485 0.504 0.550 0.568
atures (in the range of 283–313 K), at 0.6 MPa, at the stoichio- C 1.179 1.186 1.198 1.208
metric initial molar ratio of reactants methanol/acetaldehyde D 1.289 1.318 1.344 1.369
(rA/B = 2.0). All experiments lasted long enough (3 h) for the K 4.252 4.069 3.937 3.829
reaction to reach equilibrium. The experimental equilibrium
Keq = Kx .K 27.33 24.40 18.35 13.57
compositions expressed in terms of molar fractions for several
temperatures and the respective activity coefficients evaluated Experimental molar fractions and respective equilibrium constants Kx ; activity
by the UNIFAC method (Fredeslund et al., 1977) are given in coefficients and respective equilibrium constants K ; and thermodynamic
Table 2, together with the equilibrium constants expressed in equilibrium constants Keq = Kx .K .
terms of molar fractions (Kx ), activity coefficients (K ) and
activities (Keq = Kx .K ). The values of the thermodynamic 0.80
equilibrium constant Keq found in this work are similar to those
values obtained from thermodynamic experiments using Am- 0.70
Acetaldehyde Conversion

berlyst 15 as catalyst (Gandi et al., 2006). The dependency of 0.60


equilibrium constant on temperature was estimated by fitting
the experimental values by plotting ln Keq vs. 1/T (K): 0.50

0.40
2103.3
ln Keq = − 4.0595. (1) 283 K
T (K) 0.30
293 K
0.20 303 K
3.2. Reaction kinetics
0.10 313 K

The acetalization reaction between methanol and acetalde- 0.00


hyde to produce reversibly dimethylacetal (DME) and water 0 50 100 150 200 250
occurs according to Time (min)

H+ Fig. 2. Effect of temperature on conversion of acetaldehyde as function of


2CH3 OH + CH3 CHO ←→ CH3 CH(OCH3 )2 + H2 O.
time: soa = 600 rpm, P = 6 atm, V = 600 mL, wcat = 0.50 g, rA/B = 2.0.

The kinetic data consist of 14 batch experiments, where the


effect of stirring speed, temperature, initial molar ratio of re- The influence of initial molar ratio of reactants on the time
actants, catalyst loading and pressure were evaluated. All ex- evolution of the acetaldehyde conversion (Fig. 3) reveals that
periments were conducted at a stirrer speed of 600 rpm, which
ensure the elimination of external mass resistance. The qualita- (a) the higher the excess of methanol, the higher is the final
tive analysis of those parameters at the Smopex 101 kinetics is conversion of the acetaldehyde, since the excess of alcohol
similar to those obtained for Amberlyst 15 (Gandi et al., 2005). typically shifts the equilibrium towards the product forma-
As it can be observed in Fig. 2, where the limiting reactant tion;
acetaldehyde conversion was plotted along the reaction time, (b) the kinetic rate was not too affected by the initial molar
the rate of reaction increased with the increase of temperature ratio of reactants.
(10, 20, 30, and 40 ◦ C), even though the equilibrium conversion
decreased slightly due to the exothermic nature of reaction. The catalyst loading (Fig. 4) was varied over a range of 0.042
The initial molar ratio of reactants methanol (A) and ac- to 0.167% w/V on the basis of total initial volume of reactants,
etaldehyde (B) was varied from 1.5 to 4.0, corresponding to all maintaining all other reaction conditions. The effect of cata-
possible situations: lyst loading on the conversion of acetaldehyde is typical for all
heterogeneous reactions; the reaction rate is enhanced with the
(a) stoichiometric condition: rA/B = 2.0; catalyst loading increase since the total number of acidic sites
(b) excess of methanol: rA/B = 3.0 and 4.0; available increases (Chopade and Sharma, 1997). For a certain
(c) excess of acetaldehyde: rA/B = 1.5. time it is verified that the conversion is found to increase with
910 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

1.00 Table 3
The kinetic parameter as a function of temperature
0.90
Temperature (K) kC (mol/g min)
Acetaldehyde Conversion

0.80
0.70 293.15 1.67
303.15 2.92
0.60
313.15 5.99
0.50
0.40 1.5
0.30 2.0
analogy with the homogeneous catalysis, the simple mechanism
0.20 3.0
of the above reaction for the synthesis of dimethylacetal for
0.10 4.0
homogeneous is given in three steps (Morrison and Boy, 1983;
0.00 Chopade and Sharma, 1996); the reaction between acetaldehyde
0 50 100 150 200 250 300 and methanol leads to the formation of intermediate hemiacetal
Time (min) which is not stable. The hemiacetal is then reacted with the
second methanol molecule to form the final product acetal. The
Fig. 3. Effect of initial molar ratio of reactants (methanol/acetaldehyde) on
the acetaldehyde conversion as function of time: T = 293.15 K, P = 6 atm,
various steps that take place during this reaction are shown
V = 600 mL, wcat = 0.5 g. below.
Step 1: Surface reaction of methanol (A) and acetaldehyde
(B) to give hemiacetal (I1 ):
0.80
A + B ←→ I1 .
0.70
Aldehyde Conversion

0.60 Step 2: Surface reaction to obtain water (D) and intermediate


0.50 species (I2 ):

0.40 H+
0.042 % w/V I1 ←→ I2 + D (rate controlling step).
0.30
0.083 % w/V
0.20 0.125 % w/V
Step 3: Surface reaction of formation of acetal (C):
0.167 % w/V
0.10 H+
I2 + A ←→ C.
0.00
0 50 100 150 200 250
Considering this pseudo-homogeneous mechanism, the reac-
Time (min)
tion rate equation is then given by
Fig. 4. Effect of catalyst loading on acetaldehyde conversion as function of  
time: T = 293.15 K, P = 6 atm, V = 600 mL, rA/B = 2.0. aC a D
R = kC aA a B − . (2)
Keq aA

increase in the catalyst loading; however, after a certain value, This rate expression is in accordance with the Langmuir–
further increases in catalyst loading leads to a marginal increase Hinshelwood model when the adsorption constants for all
in acetaldehyde conversion. Obviously, the equilibrium com- components are negligible.
position will be the same for all cases, just the time needed to The kinetic constant (kC ) at each temperature was obtained
achieve it will be higher for low catalyst loadings. from the experimental data by optimization. The values of the
kinetic parameter kC are presented in Table 3. The temperature
3.3. Kinetic modeling dependence of the kinetic constant was fitted with the Arrhenius
equation (Fig. 5), where the activation energy obtained was of
The kinetic model based on a Langmuir–Hinshelwood rate 48.67 kJ/mol. The Arrhenius equation fitted to the experimental
expression for acetalization reactions was already derived in kinetic parameters is given below:
previous works (Silva and Rodrigues, 2001, 2006). However, 
in this case the methodology used is different because it is −5853.8
kC = 7.59 × 10 exp8
(mol/g min),
supposed that the reaction takes place on the active sites on the T (K)
surface of the catalyst due to the absence of the pores in the r 2 = 0.9914. (3)
Smopex 101 catalyst. Since the surface area of Smopex fibres is
much lower than the specific area of the available heterogeneous Fig. 6 shows the comparison between experimental and sim-
macroporous catalysts (such as Amberlyst 15), the adsorption ulated results for acetaldehyde, methanol and products (acetal
of the species at the surface can be negligible. Therefore, by and water) at 293.15 K. The average standard error between
G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918 911

the experimental and simulated molar fractions of all species 4. Fixed adsorptive reactor
is 2.0%. The kinetic model is in agreement with experimental
results for the range of initial molar ratio of reactants from 1.5 4.1. Experimental set-up
to 4.0 and at the temperature range 293.15–313.15 K.
The laboratory-scale experimental set-up (Gandi et al.,
2006) and the characteristics of the column bed are given in
Table 4. The bed was prepared by introducing glass beads in
2
the jacketed glass column together with the Smopex 101 fibres
in order to avoid building of pressure in the column due to the
1.5 compression of the bed packed only with Smopex 101 fibres.
A known amount of Smopex 101 slurry was prepared with the
Ln kc

1
Table 4
Characteristics of the fixed bed reactor
0.5
Solid weight (Smopex 101 fibres) 3.4 × 10−3 kg
Length of bed (L) 9.5 × 10−2 m
0 Internal diameter (Di ) 2.6 × 10−2 m
0.0031 0.0032 0.0033 0.0034 0.0035 Bulk density of smopex fibres (b ) 410 kg/m3
1/T Apparent density of Smopex fibres (f ) 567 kg/m3
Mass of glass beads 0.076 kg
Diameter of glass bead 0.003 m
Fig. 5. Representation of the experimental values of ln[k(mol/g min)] as
Volume occupied by glass beads 31.0 × 10−6 m3
function of 1/[T (K)] and respective linear fitting.

a b
9 10
Methanol 9 Methanol
8 T=293.15, T=293.15,
rA/B = 2.0 Products
7 rA/B = 1.50 Products 8
Acetaldehyde
Number of Moles
Number of Moles

Aldehyde 7
6
True Kinetics True Kinetics
6
5
5
4
4
3
3
2 2
1 1
0 0
0 50 100 150 200 250 0 100 200 300 400
Time (min) Time (min)

c d
12 12
Methanol Methanol
T=293.15, T=293.15,
10 rA/B = 3.0 Products 10 rA/B = 4.0 Products
Aldehyde Aldehyde
Number of Moles

Number of Moles

8 8
True Kinetics True Kinetics

6 6

4 4

2 2

0 0
0 100 200 300 400 0 50 100 150 200 250 300
Time (min) Time (min)

Fig. 6. Comparison between experimental and true kinetic model curves at T = 293.15 K, P = 6 atm, V = 600 mL, wcat = 0.50 g. (a) rA/B = 1.50; (b) rA/B = 2.0;
(c) rA/B = 3.0; (d) rA/B = 4.0.
912 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

8 lion or moles per unit volume or mass per unit volume. The
54 3 2 concentration of adsorbent on the solid is given as moles or
7 1 1 Q= 2 mL/min
mass adsorbed per unit mass or unit volume of original adsor-
2 Q= 3 mL/min
bent. Some authors estimated the mono- and multi-component
Detector Response (mV)

6 3 Q= 4 mL/min
4 Q= 5 mL/min isotherms based on the swelling ratio values of the polymeric
5 5 Q= 6 mL/min adsorbent. The swelling ratio values for Smopex 101 adsor-
bent at 20 ◦ C are 1.38 for water, 1.25 for ethanol and 1.18
4
for propanol (Lilja, 2005). More details about the swelling be-
3 haviour of Smopex 101 fibres for mono and binary systems are
also presented in Lilja (2005). In this work, the methodology is
2 based on experimental results of dynamic adsorption of binary
non-reactive mixtures to obtain binary adsorption equilibrium
1
data. Therefore, since no adsorption data for single component
0 were measured, the multi-component adsorption equilibrium
0 4 8 12 16 20 related to the fibre surface concentration was described by the
Time (min) extended Langmuir model:

Fig. 7. Tracer experiment using Blue Dextran solution. Qi Ki Cs,i


qis =  , (4)
1 + nj=1 Kj Cs,j

Table 5 where qi is the adsorbed concentration at equilibrium of


Bed porosity and Peclet number in the column species i, Qi is the adsorption capacity of species i (moles/unit
Q (ml/min)  (min) ε 2 Pe mass of fibres), Ki is the adsorption constant of species i
and Cs,i is the liquid concentration of species i at the particle
Run 1 2 6.912 0.274 13.74 198 surface.
Run 2 3 4.429 0.264 7.87 308
Run 3 4 3.481 0.276 3.15 410
Run 4 5 2.578 0.256 1.53 530 4.3. Fixed bed reactor model
Run 5 6 2.358 0.269 0.86 605
The mathematical model used to describe the dynamic be-
haviour of the fixed bed reactor, considers the following as-
methanol; the column was packed with the slurry along with the sumptions:
known weight and volume of glass beads. The porosity of the
bed prepared with Smopex 101 adsorbent and supporting inert 1. the flow pattern is described by the axial dispersed plug flow
material glass beads was determined by tracer experiments. model;
To determine the Peclet number and the bed porosity, pulse 2. external mass-transfer resistance is considered;
experiments of tracer were performed, a non-reactive substance, 3. the reaction and adsorption of species occur at the fibre
blue dextran solution, with low concentrations (5 kg/m3 ) was surface;
used to perform tracer experiments. Samples of 0.2 cm3 were 4. thermal effects are not taken into account;
injected at the inlet of the column under different flow rates, 5. constant column length and packing porosity are assumed;
and the column response was monitored at the end of the col- 6. the fibres are non-porous and the geometry is approximated
umn using a UV–VIS detector at 300 nm. The bed porosity was to infinite cylinders.
calculated from the stoichiometric time of the experimental re-
sponse curves. Calculating the second moment of the exper- The model equations consist of the following system of four
imental curves, an average Peclet number can be determined second order partial differential equations in the bulk concen-
for the range of flow rates to be used in the column. In Fig. 7, tration of the ith component, Ci , and four ordinary differential
different tracer experiments were plotted and the correspond- equations in the concentration of the ith component at the par-
ing values of bed porosity and the Peclet number are given in ticle surface, Cs,i :
Table 5. Bulk fluid mass balance to component i:
jCi jCi 2
4.2. Adsorption isotherms ε +u + f (1 − ε) Kf,i (Ci − Cs,i )
jt jz rp
 
j jxi
The adsorption isotherm is the equilibrium relationship be- = εDax CT . (5)
jz jz
tween the concentration in the fluid and the concentration in
the adsorbent particles at a given temperature and pressure. Fibres surface mass balance to component i:
Particularly for liquid adsorption, role of pressure is not that  s 
effective as in the case of gas adsorption. Liquid adsorption is 2 jqi
Kf,i (Ci − Cs,i ) = f − i R . (6)
often expressed in mass concentrations, such as parts per mil- rp jt
G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918 913

The above mass balance equations are associated with the initial where Shp = ke dp /Dm and Rep = dp u/
are, respectively,
and Danckwerts boundary conditions: the Sherwood and Reynolds numbers relative to the particle,
and Sc =
/Dm is the Schmidt number.
at t = 0,Ci = Cs,i = Ci,0 , (7) The diffusivities in multi-component liquid mixture were es-

jCi  timated by the modified Wilke–Chang equation, proposed by
z = 0, uC i − εDax = uC i,F , (8a)
jz z=0 Perkins and Geankoplis (Reid et al., 1988):

jCi  ( M)1/2 T
z = L, = 0, (8b) o
DAm = 7.4 × 10−8 , (18)
jz z=L
Vml,A
0.6

where f is fraction of solid volume due to fibres, u is the super-

n
ficial velocity, Kf is the external film mass transfer coefficient, M = x j j M j , (19)
Dax is the axial dispersion coefficient, t is the time variable, j =1
z is the axial coordinate, i is the stoichiometric coefficient of j =A
component i, f represents the apparent density of fibres, R is o is the diffusion coefficient for a dilute solute A into
where DAm
the rate given by Eq. (2), the subscripts F and 0 refer to the feed
a mixture of n components, T is the temperature, Vml,A is the
and initial states, respectively, and subscript f represents fibres.
molar volume of the solute A, xj is the molar fraction for com-
Introducing the parameter space time  = εL/u and the di-
ponent j, j is the association factor for component j, while is
mensionless variables for space  = z/L nand time  = t/, equal to 2.6 for water, 1.9 for methanol and 1.0 for unassoci-

n = C/C F,T , q̃ = q/q F,T where C F,T = i=1 CF,i and qF,T = ated components. The mixture viscosity
was predicted by the
i=1 qi (CF,1 , CF,2 , . . . , CF,n ), and the parameters, generalized corresponding state method proposed by Teja and
uL Rice (1981).
Pe = (Peclet number), (9)
εDax The system of partial differential equations represented by
Eqs. (13)–(16) are solved by the method of lines (MOL) using
kc 
Da = (Damköhler number), (10) orthogonal collocation in finite elements to discretize the axial
qF,T coordinate, with B-splines as base functions through numerical
1−ε 2 package PDECOL. Fifty subintervals for spatial discretization
Nf,i = f Kf,i  along the z-axis were used, with two internal collocation points
ε rp
(number of film mass transfer units), (11) in each subinterval, resulting in 200/400 time-dependent ordi-
nary differential equations for adsorption/reaction simulations.
(1 − ε)f f qF,T For all simulations a tolerance value EPS equal to 10−7 was
=
εCF,T fixed.
(mass capacity factor at feed conditions). (12)
The above model Eqs. (5) and (6), respectively, become as 4.4. Experimental results and discussions
follows:
 4.4.1. Binary adsorption experiments
jC̃i jC̃i 1 j j2 xi The breakthrough curves of methanol, acetal (DME) and wa-
+ + Nf,i (C̃i − C̃s,i ) = C̃T 2 , (13)
j j P e j j ter were measured in the absence of reaction. The adsorbent
was saturated with a certain component A, and then the feed
jq̃is with pure component B was fed to the adsorbent on stepwise
Nf,i (C̃i − C̃s,i ) = − i
j concentration of the feed solution. The adsorption parameters
× Da(as,A as,B − as,C as,D /(Keq as,A )), were optimized by minimizing the difference between the ex-
(14) perimental and simulated stoichiometric times of breakthrough
curves. The stoichiometric time can be determined from the
with initial and boundary conditions
mass balance over the adsorbent bed in the column as shown
 = 0, Ci = Cs,i = C0,i , (15) below:
  
1 jCi  L ∗ q
 = 0, Ci − = Ci,F , (16a) tf = ε + f , (20)
P e j =0 u C

jCi  where the parameter ∗f is defined as
 = 1, = 0. (16b)
j =1 mf
∗f = (21)
The external mass transfer coefficient was estimated by the Vb
Wilson and Geankoplis correlation (Ruthven, 1984):
and mf is the mass of fibres and Vb is the bed volume in the
1.09 column. The adsorption amount q is expressed in number of
Shp = (Rep Sc)0.33 , 0.0015 < Rep < 55, (17)
ε moles adsorbed per mass of fibres (mol/kgfibre ).
914 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

30 60

Methanol
25 50

Outlet Conc. (mol/L)


Acetal
Outlet Conc. (mol/L)

Theoretical 40
20

30 Methanol
15
Water
20
10 Theoretical

10
5
0
0 0.0 5.0 10.0 15.0 20.0 25.0 30.0
0 50 100 150 Time (min)
Time (min)
60
30
50 Methanol

Outlet Conc. (mol/L)


25 Water
40
Outlet Conc. (mol/L)

Theoretical
20
30
Methanol
15
Acetal 20
10 Theoretical
10
5
0
0 50 100 150 200 250
0
0.0 10.0 20.0 30.0 40.0 Time (min)
Time (min)
Fig. 9. Breakthrough experiments, outlet concentration of methanol and wa-
ter as a function of time. (a) Water displacing methanol; Q = 4.0 mL/ min,
Fig. 8. Breakthrough experiments, outlet concentration of acetal and methanol
bottom-up flow direction; (b) methanol displacing water; Q = 4.0 mL/ min,
as a function of time. (a) Acetal displacing methanol; Q = 4.0 mL/ min,
top-down flow direction. Experimental data: () methanol; ( ) water. The-
bottom-up flow direction; (b) methanol displacing acetal; Q = 4.0 mL/ min,
oretical results: (—).
top-down flow direction. Experimental data: () methanol; () acetal. The-
oretical results: (—).

The apparent density of the fibres can be calculated by of acetal, Fig. 8(a), the errors between the experimental and
∗f theoretical amounts of acetal adsorbed and methanol displaced
mf
f = = , (22) are 3.7% and 6.4%, respectively. Comparing the experimental
f (1 − ε) Vf amount of methanol adsorbed in the experiment of Fig. 8(b)
where with the experimental amount of methanol desorbed in the
experiment of Fig. 8(a), the error obtained is 4.9%. For acetal,
q q(CF ) − q(C0 )
= . (23) the error between the amount adsorbed, Fig. 8(a), and the
C CF − C 0 amount eluted, Fig. 8(b), is 3.8%.
Experimental stoichiometric times of the breakthrough Similarly, experiments were conducted for the combination
curves were calculated from the plot of outlet concentration of methanol and water breakthroughs and the breakthrough
of column and time of outlet samples analyzed. In the break- plots for methanol and water are shown in Figs. 9(a,b);
through experiments, where component A displaces component the experimental results are in agreement with the model
B, the total amount of A necessary to saturate the column is predictions.
given by the product of volumetric flow rate and the area over Using the experimental data and theoretical parameters, ad-
the breakthrough curve limited by the feed concentration. The sorption of methanol, acetal (DME) and water on the adsorbent
product between the flow rate and the area under the effluent (Smopex 101) were calculated at different steps of feed solu-
curve gives the total amount of component B that was ini- tions so as to determine the adsorption isotherms. Adsorption
tially saturating the column. In Figs. 8(a, b), the concentration parameters for methanol, acetal and water were calculated with
graphs of methanol and acetal are shown. In the breakthrough the data obtained from the stepwise and binary experiments
G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918 915

Table 6
Adsorption isotherms 30

Component Q 3
(mol/mreal K (m3 /mol)
solid )
25
0.93 × 10−3 Theoretical

Outlet Conc. (mol/L)


Methanol (A) 7.0
Acetaldehyde (B) 7.0 0.38 × 10−3 20 Methanol
Acetal (C) 7.0 0.84 × 10−3
Acetal(DME)
Water (D) 7.0 0.93 × 10−3 15
Aldehyde
Water
10

of both methanol–water and acetal–methanol systems. The ad-


5
sorption parameters for acetaldehyde were determined by opti-
mization of reaction experimental data. The adsorption param-
0
eters for acetal, acetaldehyde, methanol and water are given in 0 20 40 60
Table 6. Time (min)

4.4.2. Reaction experimental procedure 30


Reaction experiments were performed by feeding continu-
ously the mixture of methanol and acetaldehyde to the fixed 25

Outlet Conc. (mol/L)


bed chromatographic reactor which was initially saturated with
methanol. As the feed mixture is less dense than the pure 20
methanol, the flow direction was operated in top-down. How- Methanol
15 Acetal(DME)
ever, hydrodynamic effects due to axial backmixing driven by
natural convection of water were noticed. The concentration Aldehyde
10
fronts moving within the column are hydrodynamically stable Water
if the component above the front is less dense than the com- 5 Theoretical
ponent below the front. Since the products are denser than the
reactants, the reaction mixture is denser than methanol (ini- 0
tially saturating the column); therefore, the reactants mixture 0 5 10 15 20 25 30
should be fed from the bottom. For DME production at 20 ◦ C, Time (min)
the reactant mixture of methanol and acetaldehyde at the sto-
Fig. 10. Concentration profiles in a fixed bed adsorptive at 20 ◦ C and
ichiometric ratio was fed to the column until the steady state Q = 3.0 mL/ min: (a) reactor initially saturated with methanol and then fed
was attained; then, the bed of Smopex 101 fibres was regen- with a methanol (16.0 mol/L) and acetaldehyde (6.3 mol/L) mixture; (b) re-
erated by methanol displacement. The outlet concentration of generation step with methanol (24.7 mol/L). Experimental data: () methanol,
the reaction mixture was represented as a function of time for () acetaldehyde, () acetal and (×) water. Simulated results: (—).
the production and regeneration steps, as shown in Figs. 10(a)
and (b), respectively.
During the production, as the reactant mixture enters the col-
umn, the acetaldehyde is adsorbed and reacts with the methanol (Cmethanol =5.9×103 mol/m3 ,Cacetaldehyde =2.0×103 mol/m3 ).
initially adsorbed on the adsorbent surface. Acetal (DME) and After this reaction zone, the Smopex 101 fibres act only as ad-
water are formed in the stoichiometric amounts. The acetaliza- sorbent, where the concentration fronts of acetal, acetaldehyde,
tion reaction continues until the surface of fibres is saturated methanol and water exhibit constant pattern behaviours, since
with the reaction mixture. The concentration of the reaction they propagate along the column at constant velocities without
mixture in the column remains constant as the steady state is changing their shapes as shown in Figs. 11(a–d).
achieved, and the reaction mixture at the steady state position When the steady state is attained, the fibres (Smopex 101)
consists of the concentrations in the outlet stream. Fig. 10(a) should be regenerated with the desorbent in order to reuse the
shows that the model predictions are in agreement with the ex- fibres repeatedly before starting the new experiment. Loaded
perimental data. In this reaction experiment, one reactive front adsorbents can be regenerated by temperature-swing, pressure-
where acetalization occurs is observed, which exits the bed be- swing adsorption process or by displacement method. During
tween 2 and 8 min. These phenomena are better understood regeneration of the fixed bed adsorber by the displacement
with the analysis of propagation of these composition fronts in- method, the adsorbate is removed from the adsorbent active
side the column. This is possible only by considering the sim- sites due to the adsorption of the displacing agent. The choice
ulated internal profiles obtained from the simulation of experi- of a suitable displacement agent depends both on the equilib-
ment represented in Fig. 10(a). Figs. 11(a–d) show the internal rium of the system and on the kinetics of the adsorption and
concentration profiles for all species at 2, 4, 5 and 6 min. Near desorption. Here methanol was used as a desorbent for regen-
the feed port of the reactor, there is a reaction zone where the eration of the bed, being fed to the column in the bottom-up
reactants are rapidly consumed till the equilibrium composition direction, Fig. 10(a).
916 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

30 6

Acetaldehyde Conc. (mol/L)


25 5
Methanol Conc. (mol/L)

20 2.0 min
4
4.0 min

15 3
6.0 min
10 2
8.0 min
8.0 min
5 1 2.0 min 4.0 min 6.0 min

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Dimensionless axial distance Dimensionless axial distance

5
6
4.5 8.0 min
4 5 8.0 min
3.5 2.0 min 4.0 min 6.0 min 2.0min 4.0 min 6.0min
4
Acetal Conc. (mol/L)

Water Conc. (mol/L)

3
2.5 3
2
1.5 2

1
1
0.5
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Dimensionless axial distance Dimensionless axial distance

Fig. 11. Internal concentration profiles of each species in fluid phase inside the column, at different times, during the reaction experiment of Fig. 10(a): (a)
methanol, (b) acetaldehyde, (c) acetal and (d) water.

5. Conclusions Notation

The synthesis of acetal (1,1-dimetoxyethane) using acetalde- a liquid-phase activity


hyde and methanol as raw material on a Smopex 101 fibre both Ci concentration of component i, mol/m3
in batch reactor and in fixed bed adsorber/reactor was stud- Cs,j concentration at the particle surface, mol/m3
ied. The thermodynamic equilibrium constant was measured in dp mean pellet diameter, m
the temperature range of 293.15–333.15 K and given as Keq = Dax axial dispersion coefficient, m2 /s
0.0173 exp[2103.3/T (K)]. The proposed kinetic law expressed o
Dj,m diffusion coefficient for a dilute solute j in mixture
in activities is based on a pseudo-homogeneous catalysis model of n components, m2 /s
and the activation energy of the reaction is 48.7 kJ/mol. Ea,C reaction activation energy, J/mol
The dynamic behaviour of the fixed bed adsorptive reac- f fraction of solid volume due to fibres
tor was studied. The multi-component Langmuir adsorption kc kinetic constant, mol/kg s
isotherms were obtained from the experiments conducted to K equilibrium adsorption constant
collect the desired adsorption data by performing dynamic Keq equilibrium reaction constant, m3 /mol
binary adsorption experiments in the absence of reaction at Kf external film mass transfer coefficient, m/s
293.15 K, in a laboratory scale column. The developed mathe- Kx constant based on molar fractions
matical model for the adsorptive reactor, which used the ther- K constant based on activity coefficients
modynamic, kinetic and adsorption data measured in this work, m mass,kg
gives a good prediction of the experimental results of the fixed M molecular weight, mol/kg
bed adsorptive reactor for both production and regeneration n number of moles in the liquid phase, mol
steps. qi adsorbed concentration, mol/kg
G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918 917

Q adsorption capacity, mol/kg Chopade, S.P., Sharma, M.M., 1997. Reaction of ethanol and formaldehyde:
rp particle radius, m use of versatile cation-exchange resins as catalyst in batch reactors and
reactive distillation columns. Reactive and Functional Polymers 32, 53–64.
rA/B initial molar ratio of reactants
Doraiswamy, L.K., Sharma, M.M., 1987. Heterogeneous Reactions: Analysis,
R rate of reaction, mol/kg s Examples, and Reactor Design. Wiley-Interscience, New York.
t time, s Dupont, P., Védrine, J.C., Paumard, E., Hecquet, G., Lefebvre, F., 1995.
T temperature, K Heteropolyacids supported on activated carbon as catalysts for the
u superficial velocity, m/s esterification of acrylic acid by butanol. Applied Catalysis A 129,
217–227.
V solution volume, m3
Fedriani, A., Gonzalez-Velasco, J.R., Lopez-Fonseca, R., Gutierrez-Ortiz,
Vml molar volume of component i, m3 /mol M.A., 2000. Kinetics of 1,1-diethoxyethane synthesis catalysed by USY
wcat mass of catalyst, kg zeolites. CHISA 2000, A7.7 12.30 presentation.
x molar fraction Fredeslund, A., Gmehling, J., Rasmussen, P., 1977. Vapor–Liquid Equilibrium
X conversion of the limiting reactant Using UNIFAC. Amsterdam, Elsevier.
Gandi, G.K., Silva, V.M.T.M., Rodrigues, A.E., 2005. Process development
z axial coordinate, m
for the dimethylacetal synthesis: thermodynamics and reaction kinetics.
Industrial Engineering and Chemistry Research 44, 7287–7297.
Greek letters Gandi, G.K., Silva, V.M.T.M., Rodrigues, A.E., 2006. Synthesis of 1,1-
dimethoxyethane in fixed bed reactor. Industrial Engineering and Chemistry
 activity coefficient Research 45, 2032–2039.
Gupta, B.B.G., 1987. Alkoxyalkylation of phenol. US Patent 4, 694,111.
ε porosity
Kariner, R.S., Krause, A.O.I., 2001. Kinetic model for the etherfication

mixture viscosity of 2,4,4-trimethyl-1-pentene abd 2,4,4-trimethyl-2-pentene with methanol.


 stoichiometric coefficient Industrial Engineering Chemistry Research 40, 6073–6080.
p particle density, kg/m3 Kariner, R.S., Krause, A.O.I., Ekman, K., Sundell, M., Peltonen, R., 2000.
f apparent density of fibres, kg/m3 Etherification over a novel acid catalyst. Studies in Surface Science and
Catalysis 130D, 3411–3416.
association factor Kekre, S.Y., Gopala Rao, M., 1969. Kinetics of the liquid phase esterification
of n-butanol and acetic acid catalyzed by cation exchange resins. Indian
Subscripts Chemical Engineer 4, 115.
Lacaze-Dufaure, C., Moulougui, Z., 2000. Catalysed or uncatalysed
0 initial value esterification reaction of oleic acid with 2-ethyl hexanol. Applied Catalysis
A 204, 223–227.
A methanol Lilja, J., 2005. A fibrous polymer-supported sulphonic acid catalyst in
B acetaldehyde esterification processes. Ph.D. Thesis, Faculty of Engineering, Åbo
C acetal (DME) Akademi University.
D water Lilja, J., Aumo, J., Salmi, T., Murzin, D.Yu., Maki-Arvela, P., Sundell, M.,
eq equilibrium Ekman, K., Peltonen, R., Vainio, H., 2002a. Kinetics of esterification of
propunoic acid with methanol over a fibrous polymer-supported sulphonic
exp experimental data acid catalyst. Applied Catalysis A 228, 253–267.
f fibres Lilja, J., Murzin, D.Yu., Salmi, T., Aumo, J., Maki-Arvela, P., Sundell,
F relative to feed M., 2002b. Esterification of different acids over heterogeneous and
i relative to component i homogeneous catalysts and correlation with the Taft equation. Journal of
s surface condition Molecular Catalysis A: Chemical 182–183, 555–563.
Ma, Y., Wang, Q.L., Yan, H., Ji, X., Qiu, Q., 1996. Zeolite-catalyzed
esterification I. Synthesis of acetates, benzoates and phthalates. Applied
Acknowledgements Catalysis A 139, 51–57.
Maki-Arvela, P., Salmi, T., Sundell, M., Ekman, K., Peltonen, R., Lehtonen, J.,
1999. Comparison of polyvinylbenzene and polyolefin supported sulphonic
Ganesh K. Gandi gratefully acknowledges the PhD Grant
acid catalysts in the esterification of acetic acid. Applied Catalysis A 184,
(Research Fellowship: SFRH/BD/2782/2000) from Fundação 25–32.
para a Ciência e a Tecnologia (FCT, Portugal). This work was Morrison, R., Boyd, R., 1983. Organic Chemistry, fourth ed., Allyn and
financially supported by the project POCTI/EQU/40695/2002 Bacon, London.
(FCT, Portugal). Pääkkönen, P.K., Krause, A.O.I., 2003. Comparative study of TAME synthesis
on ion-exchange resin beads and fibrous ion-exchange catalyst. Reactive
References and Functional Polymers 55, 139–150.
Ramalinga, K., Vijayalakshmi, R., Kaimal, T.N.B., 2002. A mild and efficient
Aizawa, T., Nakamura, H., Wakabayashi, K., Kudo, T., Hasegawa, H., 1994. method for esterification and transesterification catalyzed by iodine.
Process for producing acetaldehyde dimethylacetal. US Patent 5,362,918. Tetrahedron Letters 43, 879–882.
Aumo, J., Lilja, J., Maki-Arvela, P., Salmi, T., Sundell, M., Vainio, H., Reid, C.R., Prausnitz, J.M., Poling, B.E., 1988. The Properties of Gases &
Murzin, D.Yu., 2002. Hydrogenation of citral over a polymer fiber catalyst. Liquids. fourth ed. New York, McGraw-Hill.
Catalysis Letters 84, 219–224. Roy, R., Bhatia, S., 1987. Kinetics of esterification of benzyl alcohol with
Bauer, K., Garbe, D., Surburg, H., 1990. Common Fragrance and Flavor acetic acid catalysed by cation-exchanged resin (Amberlyst-15). Journal
Materials. Wiley-VCH, Weinheim. of Chemical Technology and Biotechnology 37 (1), 1–10.
Chopade, S.P., Sharma, M.M., 1996. Esterification of formic acid, acrylic Ruthven, D.M., 1984. Principles of Adsorption and Adsorption Processes.
acid and methacrylic acid with cyclohexene in batch and distillation Wiley, New York.
column reactors: ion-exchange resins as catalysts. Reactive and Functional Schwegler, M.A., van Bekkum, H., 1991. Heteropolyacids as catalysts for
Polymers 28, 263–278. the production of phthalate diesters. Applied Catalysis A 74, 191.
918 G.K. Gandi et al. / Chemical Engineering Science 62 (2007) 907 – 918

Silva, V.M.T.M., Rodrigues, A.E., 2001. Synthesis of diethyl acetal: Teja, A., Rice, P., 1981. Generalized corresponding states method for
thermodynamic and kinetic studies. Chemical Engineering Science 56, the viscosities of liquid mixtures. Industrial and Engineering Chemistry
1255–1263. Fundamentals 20, 77.
Silva, V.M.T.M., Rodrigues, A.E., 2006. Kinetic studies in a batch reactor Yadav, G.D., Kishnan, M.S., 1998. An ecofriendly catalytic route for the
using ion-exchange resin catalysts for oxygenates production: the role of preparation of perfumery grade methyl anthranilate from anthranilic acid
mass transfer mechanisms. Chemical Engineering Science 61, 316–331. and methanol. Organic Process Research Development 2, 86–95.
da Silva-Machado, M., Cardoso, D., Pérez-Pariente, J., Sastre, E., 2000. Yadav, G.D., Pujari, A.A., 1999. Kinetics of acetalization of perfumery
Esterification of lauric acid with glycerol using modified zeolite beta as aldehydes with alkanols over solid acid catalysts. Canadian Journal of
catalyst. Studies in Surface Science and Catalysis 130D, 3417–3422. Chemical Engineering 77, 489–496.
Tanaka, K., Yoshikawa, R., Ying, C., Kita, H, Okamoto, K.I., 2002.
Application of zeolite T membrane to vapor-permeation-aided esterification
of lactic acid with ethanol. Chemical Engineering Science 57 (9),
1577–1584.

Вам также может понравиться