Anda di halaman 1dari 6

Applied Catalysis A: General 396 (2011) 1419

Contents lists available at ScienceDirect


Applied Catalysis A: General
j our nal homepage: www. el sevi er . com/ l ocat e/ apcat a
Kinetics of esterication of acetic acid with n-amyl alcohol in the presence of
Amberlyst-36
Elif des Akbay, Mehmet R. Altokka

Department of Chemical Engineering, Anadolu University, 26470, Eskisehir, Turkey


a r t i c l e i n f o
Article history:
Received 18 October 2010
Received in revised form
31 December 2010
Accepted 6 January 2011
Available online 12 January 2011
Keywords:
Esterication
Amyl acetate
Catalysis
Kinetic modeling
Amberlyst-36
a b s t r a c t
The kinetics of esterication of acetic acid with amyl alcohol has been studied in the presence of
Amberlyst-36 in its H
+
form in a batch reactor. Experimental ndings showed that the reaction is con-
trolled by chemical step rather than external and internal mass transfer steps. Experimental data well
tted the kinetic model based on homogeneous reversible reaction. Temperature dependency of the
equilibrium constant was found to be K
e
=exp(3.25820/T) which indicates that the heat of reaction is
6.81kJ/mol. It was also shown that the reaction rate can be given as
r
A
= (k
1
+k

1
C
c
)
_
C
A
C
B

C
E
C
W
K
_
where
k
1
(L mol
1
min
1
) = exp
_
19.01
9620
T
_
k

1
(min
1
) = exp
_
11.56
5990
T
_
2011 Elsevier B.V. All rights reserved.
1. Introduction
Organic esters are important ne chemicals used widely in the
manufacturing of avors, pharmaceuticals, plasticizers, and poly-
merization monomers. They are also used as emulsiers in the food
and cosmetic industries. Several synthetic routes are available for
obtaining organic esters, most of which have been briey reviewed
by Yadav and Mehta [1]. The most-used methodology for ester syn-
thesis is direct esterication of carboxylic acids with alcohols in the
presence of acid/base catalysts.
Typically, esterication reactions are very slow, requiring sev-
eral days to attain equilibrium in the absence of catalyst [2].
Despite a strong catalytic effect, homogeneous catalysts such as
sulfuric acid, hydrouoric acid and para-toluene sulfonic acid
are toxic, corrosive and often hard to remove from the prod-
ucts. Thus it is keenly desirable to develop new types of solid
acid catalysts to replace the homogeneous catalysts. Addition-
ally, they have been found to offer better selectivity towards

Corresponding author. Tel.: +90 222 321 35 50/6502; fax: +90 222 323 95 01.
E-mail address: mraltiokka@anadolu.edu.tr (M.R. Altokka).
the desired product(s) compared to homogeneous catalysts
[3,4].
Many solid catalysts, e.g., new solid acids and bases, ion-
exchange resins, zeolites and acidic clay catalysts were proposed
in literature [515]. Among them, cation-exchange resins are the
most commonly used solid acid catalysts in organic reactions
[7,8].
The effects of many ion exchange resins were investigated in
esterication reactions. For example; methyl acetate distillation in
a catalytic column [16], esterication of acetic acid with methanol
and butanol [17,18], liquid-phase esterication of propionic acid
withn-amyl alcohol [19], estericationof lactic acidwithmethanol
[20], esterication of acrylic acid with propylene glycol [21] were
investigated in the presence of Amberlyst-15. Amberlyst-35 was
used in the study of liquid-phase esterication of propionic acid
with n-butanol [22,23] and Amberlite IR-120 was employed in the
esterication of acetic acid with iso-butanol [3] and esterication
of acetic acid with iso amyl alcohol was realized in the presence of
Purolite CT-175 [24].
Amyl acetate, studiedinthis work, has beenusedinindustries as
a solvent, an extractant, a polishing agent, etc. It can be synthesized
from acetic acid and amyl alcohol via an esterication reaction in
the presence of an acid catalyst [25].
0926-860X/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2011.01.013
E.. Akbay, M.R. Altokka / Applied Catalysis A: General 396 (2011) 1419 15
Nomenclature
C concentration, mol/L; catalyst
X conversion
t time, min
r reaction rate, mol/L min
T absolute temperature, K
k reaction rate constant, L/mol min
K equilibrium constant
E activation energy, kJ/mol
M C
B0
/C
A0
Subscripts
A, B, E, W and C acetic acid, n-amyl alcohol, amyl acetate,
water and catalyst, respectively
e equilibrium condition
o initial condition
Literature has shown that the heterogeneous esterication of
acetic acid with amyl alcohol has scarcely been studied. Lee et al.
[25] studied the kinetic behavior of the esterication of acetic
acid with amyl alcohol over Dowex 50Wx8-100. The experiments
were conducted in a xed-bed reactor at temperatures ranging
from 323 to 393K and at molar ratios of amyl alcohol to acetic
acid from 1 to 10. They found that the reaction is endother-
mic with 5.43kJ/mol within the working temperatures mentioned
above. They have also shown that a water phase (the second
liquid phase) appeared in the reacting mixture when the feed
composition was close to stoichiometric. The modied Langmuir-
Hinshelwood model was used to represent the kinetic behavior of
the reactionover wide ranges of temperature andfeedcomposition
[25].
The other work related to the esterication of acetic acid with
amyl alcohol was carried at in the presence of Amberlyst-15.
Homogenous kinetic model was proposedto represent the reaction
progression. The activation energy of the reaction was reported to
be 6.460kJ/mol [26].
In another study, the kinetics of heterogeneous catalyzed ester-
ication of acetic acid with amyl alcohol was investigated by
Palani et al. [27]. In the study, Al-MCM-41 molecular sieves with
Si/Al ratios of 25, 50, 75 and 100 were synthesized. Their cat-
alytic activity was tested in the vapor phase esterication between
150 and 250

C. They showed that the activity of the catalysts


followed the order Al-MCM-41(100) >Al-MCM-41(75) >Al-MCM-
41(50) >Al-MCM-41(25) [27].
Inthe present work, Amberlyst-15, Amberlyst-36andAmberlite
IR-120 were tested as the heterogeneous catalysts for the ester-
ication of acetic acid with n-amyl alcohol. Their activities were
determined. Thereafter, the kinetic study of esterication was con-
ducted in a batch reactor in the presence of the catalyst with the
highest activity among them.
2. Experimental
2.1. Materials
Acetic acid (>99.5%) was obtained from Merck. Amyl alcohol
(>99%) as one of the reactants and dioxane as the solvent were sup-
plied by SigmaAldrich. The catalysts and their properties stated
by manufacture are given in Table 1.
2.2. Apparatus
The reactor consisted of a two-necked spherical Pyrex ask of
500ml capacity tted with a spiral coil condenser and a sample
device. The temperature was controlled within 0.1K by circulat-
ing water from a thermostat into the water jacket of the reactor.
Mixing of the reaction uid was performed by a magnetic stirrer.
2.3. Experimental procedure
The catalyst was kept ina waterdioxane solutioncontaining 2%
water, which is the amount of water expected in the reaction mix-
ture, for a night before using in the experiment. Thus the swollen
effect of the catalyst was assumed to be eliminated during the reac-
tion [3].
In a typical run, dioxane as the solvent and one of the reactants
were placed in the reactor. A known amount of catalyst was added
and the reactor contents were well mixed. After a steady value of
desired temperature was attained, the second reactant was added
and this was taken as zero time for the run. Two milliliters of liquid
sample were withdrawn from the reactor at regular intervals for
analysis and immediately transferred to a crucible in an ice bath in
order to ensure that no further reaction took place.
2.4. Analysis
Acetic acid, in the reaction mixture, was determined by titration
with 0.1N standard sodium hydroxide solutions with phenolph-
thalein as an indicator. The water content in the reaction mixture
was measured by Karl Fischer titration (Metrohm KF-784). After
verifying that the measured water content corresponds to the
calculatedvalues basedonstoichiometric equation, it has beencon-
cluded that there is no by product formation. Thus, the analysis was
carried out by the determination of acetic acid only and the other
components of the mixture were calculated from stoichiometric
mass balance equation.
3. Results and discussion
3.1. Effect of parameters
The effects of parameters; such as catalyst type and loading,
external and internal mass transfer resistances, temperature and
reactant mole ratioonthe reactionrate for estericationwere stud-
ied.
Table 1
Properties of catalysts reported by manufacturer.
Amberlyst-15 Amberlyst-36 Amberlite IR-120
Manufacturer Rohm & Haas Co. Rohm & Haas Co. Rohm & Haas Co.
Matrix Styrene-divinylbenzene (macroreticular) Styrene-divinylbenzene (macroreticular) Styrene-divinylbenzene (Gel Type)
Standard ionic form H
+
H
+
H
+
Surface area (m
2
/g) 53 33
% Moisture 48 55 45
Particle size (mm) 0.3550.850 0.4250.850 0.3000.830
Cation exchange capacity (mmol/g) 4.7 5.4 4.4
Maximum operating temperature (K) 393 423 393
16 E.. Akbay, M.R. Altokka / Applied Catalysis A: General 396 (2011) 1419
0
0.1
0.2
0.3
0.4
0.5
600 500 400 300 200 100 0
X
A
t (min)
Amberlyst-15
Amberlite IR-120
Amberlyst-36
Fig. 1. Comparison of catalyst type on the reaction rate at 343K, M=1 with
C
A0
=1.94mol/L and 0.26mol H
+
/L in equivalent for each catalyst type.
3.1.1. Catalyst type
In the experiment, three types of ion exchange resins;
Amberlyst-15, Amberlyst-36, and Amberlite IR-120 have been
tested. Experiments were carriedout at 343KandM=1, inthe pres-
ence of the catalyst containing the same amount of ion exchange
capacity of 0.26mol H
+
/L in equivalent for each type. Results are
given in Fig. 1. Product selectivity of the catalysts is not consid-
ered in this study since it is assumed that there is no by products
formation. As seen from Fig. 1 that Amberlyst-36 accelerates the
reaction rate slightly more than the others since the higher conver-
sions were obtained at the same time. Therefore, the kinetic model
was developed in the presence of Amberlyt-36 in this study.
3.1.2. Mass transfer resistances
To investigate the external mass transfer resistance the exper-
iment were carried out at 300, 500 and 700rpm under the same
reaction conditions of 343K temperature, M=1, and catalyst load-
ing of 0.26mol H
+
/L in equivalent. Results are given in Fig. 2. As
seen in Fig. 2, the reaction rate is not signicantly differed by vary-
ing the stirring speed. Particularly above 500rpm, the differences
canbeconsideredtobenegligible. Thereafter theexperiments were
realized at constant stirring speed of 600rpm. Fromthis analysis it
was concludedthat the reactionis not external diffusioncontrolled.
0
0.1
0.2
0.3
0.4
0.5
600 500 400 300 200 100 0
X
A
t(min)
300 rpm
500 rpm
700 rpm
Fig. 2. Effect of stirring speed on the reaction rate at 343K, M=1 with
C
A0
=1.94mol/L and Amberlyst-36 loading of 0.26mol H
+
/L in equivalent.
0
0.1
0.2
0.3
0.4
0.5
600 500 400 300 200 100 0
X
A
t(min)
600-850 m
425-600 m
Fig. 3. Effect of catalyst particle sizes on the reaction rate at 343K, M=1 with
C
A0
=1.94mol/L and Amberlyst-36 loading of 0.26mol H
+
/L in equivalent.
The reactionwas also conductedinthe presence of different size
of the same amount of catalyst under the constant reaction condi-
tions of 343K temperature, M=1, and catalyst loading of 0.26mol
H
+
/L in equivalent to investigate the internal mass transfer resis-
tance. Results obtained in the presence of particles of size range of
425600m and 600850m are given in Fig. 3. As seen in Fig. 3,
the reaction rate is almost independent of particle size, indicat-
ing the reaction is not controlled by internal diffusion. Thereafter
experiments were realized in the presence of Amberlyst-36 whose
size is given at present in Table 1.
These results conrm the literature reports claiming that the
inuence of external and internal diffusion can be neglected for
most of the reactions catalyzed by the Amberlyst series resins
[2832].
3.1.3. Effect of temperature
Esterication reaction was realized at different temperatures of
333, 348 and 358K under the constant reaction conditions of M=1,
andcatalyst loadingof 0.26mol H
+
/Linequivalent. Results aregiven
inFig. 4. As seeninFig. 4, the reactionrate strongly depends ontem-
perature, indicating the reaction is controlled by chemical steps.
Depending on this fact, the kinetic model will be developed.
0
0.1
0.2
0.3
0.4
0.5
500 400 300 200 100 0
X
A
t(min)
333 K
348 K
358 K
Fig. 4. Effect of temperature on the reaction rate at M=1 with C
A0
=1.94mol/L and
Amberlyst-36 loading of 0.26mol H
+
/L in equivalent.
E.. Akbay, M.R. Altokka / Applied Catalysis A: General 396 (2011) 1419 17
R = 0.998
0
0.001
0.002
0.003
0.004
0.005
0.006
0.7 0.6 0.5 0.4 0.3 0.2 0.1 0
-
r
A
0
(
m
o
l
/
L
.
m
i
n
)
C
cat
(mol/L)
Fig. 5. Effect of Amberlyst-36 loading on the initial reaction rate at 343K and M=1
with C
A0
=1.94mol/L.
3.1.4. Effect of catalyst loading
Experiments were carried out in the absence of catalyst and as
well as in the presence of catalyst of 0.12, 0.26, 0.40, and 0.65mol
H
+
/L in equivalent under the constant reaction conditions of 343K
temperature and M=1 with C
A0
=1.94mol/L. The reaction rate, up
to about 10% conversion, can be considered to be a linear function
of time. Therefore the initial reaction rate can be safely calculated
by Eq. (1).
r
A,O
=
C
A,O
X
A
t
(1)
Applying Eq. (1) to the experimental data obtained at the con-
version not higher than 10% percent, the initial reaction rate was
calculated. A plot of the initial reaction rate vs. catalyst loading
is given in Fig. 5. As seen in Fig. 5, the reaction rate, as expected,
is increasing linearly with catalyst loading since the active surface
area is proportional totheamount of catalyst. Thepoint determined
by the intersection of the line with ordinate is in good agreement
with the uncatalyst reaction rate obtained experimentally at given
parameters. The mathematical expression relating the initial reac-
tionratetothecatalyst loadingcanbederivedfromFig. 5as follows:
r
A,O
(mol L
1
min
1
) = 0.0009 +0.0062C
cat
(mol L
1
) (2)
It should also be kept in mind that Eq. (2) is valid only at the
givenparameters of 343KandM=1withC
A0
=1.942mol/L at which
the experiments were performed. However, it does not alter the
conclusion that the general reaction rate increases linearly with
the catalyst loading.
3.1.5. Effect of initial reactants mole ratio
Initial reaction rate was determined by altering the concentra-
tion of the component under investigation while keeping that of
the others constant at temperature of 343K and catalyst loading of
0.26mol H
+
/Linequivalent. Theresults areshowninFig. 6. Theordi-
nate values of these gures were obtained fromEq. (1) by using the
experimental data. It is evident fromthese gures that the adsorp-
tions of acetic acid and n-amyl alcohol are negligible on the catalyst
since the initial reaction rate is increasing linearly, rather than
producing a gure of plateau, by increasing their corresponding
concentrations [33].
3.2. Determination of the equilibrium constant
The apparent equilibrium constant of the reaction can be found
from the equilibrium conversions, determined experimentally, as
0
0.002
0.004
0.006
0.008
0.01
10 8 6 4 2 0
-
r
A
0
(
m
o
l
/
L
.
m
i
n
)
C
O
(mol/L)
n-Amyl Alcohol
Acec Acid
Fig. 6. Effect of reactant concentration on the initial reaction rate at 343K and
Amberlyst-36 loading of 0.26mol H
+
/L in equivalent.
follows:
k
e
=
X
2
A,e
(1 X
A,e
)(MX
A,e
)
(3)
In Eq. (3) the equilibrium constant is stated in terms of concen-
tration instead of activities since activity can be dened as a
i
=
i
C
i
and the coefcient
i
can readily be combined with constant.
The reaction was carried out up to the equilibrium conversion
at temperatures of 333, 348 and 358K. Temperature dependency
of the equilibrium constant was also determined by applying vant
Hoff equation. The result is shown in Fig. 7.
From Fig. 7, the equilibrium constant was found to be
K
e
= exp
_
3.25
820
T
_
(4)
where T is absolute temperature in Kelvin.
The heat of reaction, assuming independent of temperature
ranging from 333K to 358K, calculated to be, H
r
=0.820
8.314=6.81kJ/mol since the heat of reaction calculated from
equilibrium constant is identical regardless of the nature of the
equilibrium constant depending on either activities or concentra-
tions.
This value is in a good agreement with the literature nding of
5.43 and 6.46kJ/mol [25,26].
R = 0.996
0.7
0.8
0.9
1
0.0031 0.003 0.0029 0.0028 0.0027
l
n
(
K
)
1/T (1/K)
Fig. 7. Temperature dependency of the equilibrium constant.
18 E.. Akbay, M.R. Altokka / Applied Catalysis A: General 396 (2011) 1419
3.3. Kinetic modeling
The general reaction stoichiometry and corresponding reac-
tion rate expression, in the form of LangmuirHinshelwood
HougenWatson (LHHW) model with surface reaction is the deter-
mining step, can be written as;
CH
3
COOH+
A
CH
5
H
11
OH
B
C
5
H
11
COOCH
3
E
+H
2
O
W
(5)
r
A
=
k
_
C
A
C
B

_
C
E
C
W
/K
_
(1 +

K
i
C
i
)
2
(6)
where K
i
and C
i
adsorption equilibrium constant and bulk concen-
tration of species i, respectively.
As stated in Section 3.1.5, acetic acid and amyl alcohol are
weakly adsorbed on the catalyst surface. On the other hand,
it was concluded that the other species in the reaction mix-
ture are also not adsorbed appreciably on the surface of catalyst
since homogeneous reversible reaction model well tted to the
experimental data. Thus, the denomination in Eq. (6) approaches
unity.
Then the general reaction stoichiometry and corresponding
reaction rate expression can be given as:
A +B
k
1

k
2
E +W uncatalyzed (7a)
A +B +C
k
1

k
2
E +W+C catalyzed (7b)
r
A
= k
1,obs
_
C
A
C
B

C
E
C
W
K
_
(8)
where C is catalyst and, since both the catalyzed and uncatalyzed
reactions aretakingplacesimultaneouslyinthepresenceof catalyst
k
1,obs
is given by
k
1,obs
= k
1
+k

1
C
c
(9)
Eq. (8) can be rearranged in terms of conversion, assuming that
there is no product initially, as
r
A
= C
A,O
dX
A
dt
= k
1,obs
C
2
A,O
_
(1 X
A
)(MX
A
)
X
2
A
K
_
(10)
If Eq. (10) is integrated and rearranged in its linear formthe follow-
ing equation will be obtained:
ln
__
1 +M+a
2
2a
1
X
A
1 +Ma
2
2a
1
X
A
__
1 +Ma
2
1 +M+a
2
__
= a
2
k
1,obs
C
A,o
t (11)
where
a
1
= 1 K
1
e
a
2
=
_
(M+1)
2
4a
1
M

1/2
and the other symbols in equation are listed in nomenclature.
A plot of the left hand side of Eq. (11), obtained experimen-
tally up to about 50% conversion corresponding to the 85% of
the equilibrium conversion of acetic acid, versus time is given in
Fig. 8. Obtaining straight lines in Fig. 8 is another verication of the
model. Forward reaction rate constant, k
1,obs
was found from the
slope of lines in Fig. 8, in the presence of the catalyst of 0.26mol
H
+
/L in equivalent at different temperatures. Results are given in
Table 2.
Applying the Arrhenius equation to the values in Table 2, the
temperature dependency of the rate constant, k
1,obs
as well as its
corresponding activation energy at the catalyst loading of 0.26mol
Fig. 8. Adopted Eq. (11) to experimental data.
Table 2
The values of the forward rate constant, k
1,obs
at different temperatures with the
catalyst loading of 0.26mol H
+
/L in equivalent.
T (K) 333 348 358
k
1,obs
(10
3
) (L mol
1
min
1
) 0.456 1.224 1.746
Table 3
The values of the forward rate constant, k
1
at different temperatures for uncatalyzed
reaction.
T (K) 333 348 358
k
1
(10
3
) (L mol
1
min
1
) 0.049 0.201 0.358
H
+
/L in equivalent were found to be;
k
1,obs
(L mol
1
min
1
) = exp
_
11.92
6520
T
_
,
E = 54.21kJ/mol (12)
where T is absolute temperature in K.
The experiments were also performed to determine the uncat-
alyzed reaction rate constant k
1
at temperatures of 333, 348 and
358K and M=1 with C
AO
=1.96mol/L. Under these conditions, to
obtain the initial reaction rate, Eq. (10) can be reduced to be
r
A,O
= k
1
C
2
A,O
M (13)
Combining of Eq. (1) with Eq. (13), k
1
was calculated at different
temperatures. Results are given in Table 3.
Applying the Arrhenius equation to the values in Table 3, the
temperature dependency of the rate constant k
1
as well as its cor-
responding activation energy were found to be;
k
1
(L mol
1
min
1
) = exp
_
19.01
9620
T
_
, E = 79.98kJ/mol
(14)
Thus, k

1
canbedeterminedfromEq. (9) byusingthevalues given
in Tables 2 and 3 since C
C
is known to be 0.26mol H
+
/L. Results are
given in Table 4.
Table 4
The values of the forward rate constant, k
1

at different temperatures.
T (K) 333 348 358
k

1
(10
3
) (L
2
mol
2
min
1
) 1.565 3.935 5.338
E.. Akbay, M.R. Altokka / Applied Catalysis A: General 396 (2011) 1419 19
Applying the Arrhenius equation to the values in Table 4, the
temperature dependency of the rate constant, k

1
as well as its cor-
responding activation energy were found to be;
k

1
= (min
1
) = exp
_
11.56 5990
T
_
, E = 49.80kJ/mol (15)
The reversed reaction rate constants k
2
and k

2
can easily be
calculated from equilibrium constant given in Eq. (4).
4. Conclusions
Three types of catalysts, Amberlyst-15, Amberlyst-36, and
Amberlite IR-120 have been tested in the esterication of acetic
acid with amyl alcohol. Preliminary investigation showed that
Amberlyst-36 has slightly higher activities than the others. There-
fore kinetic model was developed in the presence of Amberlyst-36
in its H
+
form. Experimental ndings showed that the reaction is
controlled by chemical step rather than external and internal mass
transfer steps.
Experimental data well tted the kinetic model based on
homogeneous reversible reaction. Temperature dependency of the
equilibrium constant was found to be
K
e
= exp
_
3.25
820
T
_
from which the heat of reaction was calculated as
6.81kJ/mol.Kinetics model, based on experimental data, was
also developed. It was shown that the reaction rate can be given as
r
A
=
_
k
1
+k

1
C
c
_
_
C
A
C
B

C
E
C
W
K
_
where
k
1
(L mol
1
min
1
) = exp
_
19.01
9620
T
_
k

1
(L
2
mol
2
min
1
) = exp
_
11.56
5990
T
_
References
[1] G.D. Yadav, P.H. Mehta, Ind. Eng. Chem. Res. 33 (1994) 21982208.
[2] J. Lilja, J. Aumo, T. Salmi, D. Murzin, P. Maki-Arvela, M. Sundell, K. Ekman, R.
Peltonen, H. Vainio, Appl. Catal. A: Gen. 228 (2002) 253267.
[3] M.R. Altokka, A. C tak, Appl. Catal. A: Gen. 239 (2003) 141148.
[4] M.R. Altokka, H.L. Hosgn, Ind. Eng. Chem. Res. 46 (2007) 10581062.
[5] K. Tanabe, M. Misono, Y. Ono, H. Hattori, New Solid Acid Bases, Kodan-
sha/Elsevier Science, Tokyo/Amsterdam, 1989.
[6] A.E.R.S. Khder, Appl. Catal. A: Gen. 343 (12) (2008) 109116.
[7] A. Charkrabati, M.M. Sharma, React. Polym. 20 (1993) 145.
[8] M.M. Sharma, React. Funct. Polym. 26 (1995) 323.
[9] S.B. Valdeilson, C.L. Ivoneide, F.A.C. Garcia, S.C.L. Dias, J.A. Dias, Catal. Today
133135 (2008) 106112.
[10] T.L. Marker, G.A. Funck, T. Barger, U. Hammershaimb, US Patent, 5504258,
(1996).
[11] D.E. Hendriksen, J.R. Lattner, M.J.G. Janssen, US Patent, 6002057, (1999).
[12] R. Chitnis, M.M. Sharma, React. Funct. Polym. 32 (1997) 93115.
[13] J.T. Kloprogge, J. Porous Mater. 5 (1998) 541.
[14] C.S.M. Pereira, S.P. Pinho, V.M.T.M. Silva, A.E. Rodrigues, Ind. Eng. Chem. Res. 47
(2008) 14531463.
[15] V.T.M.M. Silva, A.E. Rodrigues, Chem. Eng. Sci. 61 (2006) 316331.
[16] Y. Fuchigami, J. Chem. Eng. Jpn. 23 (1990) 354358.
[17] Z.P. Xu, K.T. Chuang, Chem. Eng. Sci. 52 (17) (1997) 30113017.
[18] J. Gangadwala, S. Mankar, S.M. Mahajani, A. Kienle, E. Stein, Ind. Eng. Chem. Res.
42 (2003) 21462155.
[19] B. Erdem, M. Cebe, Korean J. Chem. Eng. 23 (6) (2006) 896901.
[20] M.T. Sanz, R. Murga, S. Beltran, J.L. Cabezas, Ind. Eng. Chem. Res. 41 (2002)
512517.
[21] M.R. Altokka, E. des , Appl. Catal. A: Gen. 362 (2009) 115120.
[22] W.T. Liu, C.S. Tan, Ind. Eng. Chem. Res. 40 (2001) 32813286.
[23] M.J. Lee, J.Y. Chiu, H.M. Lin, Ind. Eng. Chem. Res. 41 (2002) 28822887.
[24] H.T.R. Teo, B. Saha, J. Catal. 228 (2004) 174182.
[25] M.J. Lee, H.T. Wu, H.M. Lin, Ind. Eng. Chem. Res. 39 (2000) 40944099.
[26] M.J. Lee, H.T. Wu, C.H. Kang, H.M. Lin, J. Chin. Inst. Chem. Eng. 30 (2) (1999)
117122.
[27] A. Palani, A. Pandurangan, J. Mol. Catal. A: Chem. 226 (2005) 129134.
[28] P. Delgado, M.T. Sanz, S. Beltran, Chem. Eng. J. 126 (2007) 111118.
[29] V.J. Cruz, J.F. Izquierdo, F. Cunill, J. Tejero, M. Iborra, C. Fite, R. Bringue, React.
Funct. Polym. 67 (3) (2007) 210224.
[30] M.P. Titus, M. Bausach, J. Tejero, M. Iborra, C. Fite, F. Cunill, J.F. Izquierdo, Appl.
Catal. A: Gen. 323 (2007) 3850.
[31] S.H. Ali, A. Tarakmah, S.Q. Merchant, T. Al-Sahhaf, Chem. Eng. Sci. 62 (2007)
31973217.
[32] W. Mao, X. Wang, H. Wang, H. Chang, X. Zhang, J. Han, Chem. Eng. Proc. 47
(2008) 761769.
[33] H.S. Fogler, Elements of Chemical Reaction Engineering, third ed., Prentice Hall,
New Jersey, 1999.

Anda mungkin juga menyukai