Vous êtes sur la page 1sur 8

Fuel 88 (2009) 461468

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Production of biodiesel by esterication of palmitic acid over mesoporous


aluminosilicate Al-MCM-41
Alpio C. Carmo Jr. a, Luiz K.C. de Souza a, Carlos E.F. da Costa a, E. Longo b,1, Jos R. Zamian a,
Geraldo N. da Rocha Filho a,*
a

Universidade Federal do Par (UFPA), Instituto de Cincias Exatas e Naturais, Faculdade de Qumica, Laboratrio de Catlise e Oleoqumica, 66075-110 Belm, PA, Brazil
Universidade Estadual Paulista Jlio de Mesquita Filho (UNESP), Instituto de Qumica de Araraquara, Departamento de Bioqumica e Tecnologia Qumica. 14801-907 Araraquara,
Caixa-Postal: 355 SP, Brazil
b

a r t i c l e

i n f o

Article history:
Received 7 February 2008
Received in revised form 22 September
2008
Accepted 7 October 2008
Available online 31 October 2008
Keywords:
Mesoporous
Molecular sieves
Acid catalysts biodiesel
Esterication

a b s t r a c t
Biodiesel has been obtained by esterication of palmitic acid with methanol, ethanol and isopropanol in the
presence of Al-MCM-41 mesoporous molecular sieves with Si/Al ratios of 8, 16 and 32. The catalytic acids
were synthesized at room temperature and characterized by atomic absorption spectrometry (AAS), thermal analysis (TG/DTA), X-ray diffraction (XRD), nitrogen absorption (BET/BJH), infrared spectroscopy (IR),
scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The reaction was carried
out at 130 C whilst stirring at 500 rpm, with an alcohol/acid molar ratio of 60 and 0.6 wt% catalyst for 2 h.
The alcohol reactivity follows the order methanol > ethanol > isopropanol. The catalyst Al-MCM-41 with
ratio Si/Al = 8 produced the largest conversion values for the alcohols studied. The data followed a rather
satisfactory approximation to rst-order kinetics.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
Currently, there is a great interest in the esterication reaction
because of its application to several branches of industry [1]. Organic esters are frequently used in the production of plastic derivatives, in the solvent industry, perfumery, agrochemistry and other
branches of ne chemistry [2]. One of the main products obtained
by esterication of long chain fatty acids is biodiesel, whose use
has several environmental benets [3]. Free fatty acids (FFA) poison homogeneous catalysts such as NaOH and KOH, forming soaps
and creating difculties to separate the products of the reaction
[4]. Rened oil has a low content of FFA. However, use of this kind
of oil is not economically feasible since it adds one step to the process, and also competing with food and cosmetics industries,
which also makes use of rened oils. Non rened oils, grease and
recycled oils present a high content of FFA.
Esterify FFA to alkyl esters in the presence of an acidic catalyst is
a route to improving the use of high FFA oils on biodiesel production. Esterication is normally carried out in the homogeneous
phase in the presence of acid catalysts such as sulfuric and p-toluene sulfonic acids. This pretreatment step has been successfully
demonstrated using sulfuric acid [5]. Unfortunately, use of the
homogeneous sulfuric acid catalyst adds neutralization and separa* Corresponding author. Tel.: +55 (91) 32018032.
E-mail address: narciso@ufpa.br (G.N. da Rocha Filho).
1
Tel.: +55 (16) 33220015.
0016-2361/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2008.10.007

tion steps as well as the esterication reaction to the process [4].


Esterication reaction of FFA followed by transesterication. The
use of heterogeneous catalysts in the can be an alternative to reduce
biodiesel cost.
Other types of acid catalyst such as protonated Naon resin
(H-MOR), Nb2O5  nH2O, and mordenite zeolite (H-MOR) have been
used in the esterication of oleic and estearic acids with ethanol
instead of sulfuric acid [6]. Due to corrosion and contamination
problems with the nal product because of the use of homogeneous catalysts, alternative approaches have been studied in order
to substitute them with heterogeneous materials [7,8]. Ni et al. [9]
studied the esterication of palmitic acid dissolved in commercial
sunower with methanol using SAC-13 (Naon/SiO2). Ramu et al.
[10] used a combination of the tetragonal phase of zirconia and
amorphous WO3. They reported high activity of the acid catalyst
(WO3/ZrO2) in the esterication of palmitic acid with methanol
and a good correlation was obtained between the acidity and catalytic activity, indicating a strong inuence of acidity and the presence of tetragonal zirconia.
Mesoporous molecular sieves MCM-41, materials developed by
Mobil Oil Corporation researchers [11], have attracted great interest since their discovery in the early 90s. Their physical properties
such as high specic area, volume, and pore diameter enable them
to be used as adsorbents, catalytic supports and heterogeneous
catalysts in several branches of the chemical industry such as the
production of biofuels [12], alkylation of oil fractions [13] and ne
chemistry [14,15], mainly as substituents of microporous solids

462

A.C. Carmo et al. / Fuel 88 (2009) 461468

like zeolites [16,17]. As well as having a highly ordered distribution


of mesopores, MCM-41 solids offer the possibility of incorporating
of metals into their structures, which improves their hydrothermal
stability and produces active sites, thus broadening their eld of
application [18].
Ajaikumar and Pandurangan [19] studied the inuence of the
Si/Al ratio on the esterication reaction of alkyl acids with several
alcohols and observed that the acidity increases with the incorporation of Al into the structural lattice of MCM-41, and that its concentration is directly related to the hydrophobic character of the
solid. Luque et al. [20] carried out the N-methylation reaction of
aniline in the presence of Al-MCM-41 at Si/Al ratios from 15 to
40. They observed that the acidity of Al-MCM-41 increases with
a decrease of the Si/Al ratio. Consequently, the conversion of aniline may be attributed to a greater concentration of Brnsted and
Lewis acid sites. Gokulakrishnan et al. [21] studied the esterication of acetic acid with propanol isomers using Al-MCM-41 as a
catalyst with Si/Al ratios of 30, 51, 72, and 97. The best conversion
values for the esterication of n-propanol and isopropanol were
obtained with catalyst Al-MCM-41 (Si/Al = 30); this showed the
greatest activity in comparison to other acid solids such as zeolite
HY and H mordenite.
Given the acidic nature of Al-MCM-41 mesoporous solids, the
present work aims to synthesize these materials with three Si/Al
ratios (8, 16, and 32) and later to characterize their thermal, structural, and morphological aspects and to study the catalytic production of biodiesel by the esterication of palmitic acid with
methanol, ethanol and isopropanol. The catalytic activity of the
MCM-41 catalysts was studied on the FFA removal reaction. A
technological challenge in the use of vegetable oil for biodiesel production is the use of low quality oils which contain excessive free
fatty acids. Thus, it is necessary to conduct a preliminary study on
the use mesoporous materials for the pretreatment esterication
reaction.

mobalance model PLSTA. X-ray diffractograms were obtained


with a PHILIPS diffractometer model PW 3710 equipped with a
copper anode (k Cu Ka = 1.5406 ) over a scanning interval (2h)
from 1.5 to 10. The specic surface area, and pore diameter and
volume were obtained by nitrogen adsorption analysis at 77 K using
a QUANTACHROME analyzer model NOVA 1200 and following the
BET (BrunauerEmmettTeller) and BJH (BarrettJoynerHalenda)
methods. Infrared spectroscopy analysis was performed with a
THERMO ELECTRON CORPORATION model IR 100 over the range
4004000 cm1. Scanning electron microscopy (SEM) images were
taken with a ZEISS model LEO 1430, operating at 10 kV and with
a beam current of 90 mA. Transmission electron microscopy
(TEM) images were taken with a PHILIPS model cm 200 operating
at 200 kV.
2.3. Esterication reaction
Catalysis tests were carried out in a 250-mL PAAR 4560 reactor
provided with mechanical agitation and coupled to a PAAR controller model 4853. We used 0.6 wt% catalyst at a 60/1 alcohol/acid ratio under 500 rpm agitation at 130 C for 2 h and atmosphere
pressure. At the end of each experiment, the nal acidity was
determined by titration with potassium hydroxide. To verify the
performance of the catalyst of the high free fatty acid oil esterication, palm oil modied with 17% of oleic acid was used in two temperatures, maintaining the parameters as like the ones in the
model reaction with palmitic acid.
The conversion of FFA was calculated by the formula:

xffa %

ai  at
100;
ai

where ai is the initial acidity of the mixture and at is the acidity at a


t time.
3. Results and discussion

2. Experimental

3.1. Characterization

2.1. Synthesis of Al-MCM-41

The amounts of Al in Al-MCM-41 catalysts for Si/Al ratios of 8,


16, and 32 are presented in Table 1. These masses were anticipated
in the initial calculations by varying the amount of the aluminum
source since the number of moles of the silicon source (TEOS) was
kept constant at approximately 1.255. AAS results show that the
amounts of aluminum obtained did not depart signicantly from
the theoretical amounts.
Thermal analysis results (TG/DTA) of Al-MCM-41 catalysts
without thermal treatment are presented in Fig. 1.
According to the TG curves of the Al-MCM-41 samples not thermally pre-treated, mass losses of 4252% occur between 25 and
850 C. These curves reveal the appearance of three distinct mass
loss regions. The rst loss (7.5%), between 30 and 190 C, is attributed to the removal of supercial water molecules or water from
the solid pores, which is conrmed by the endothermic peak that
appears in this temperature interval in the DTA curves. The second
mass loss (35.5%) occurs between 190 and 400 C followed by
an exothermic peak with high intensity, characteristic of oxidation

Catalysts Al-MCM-41 with Si/Al ratios of 8, 16, and 32 were synthesized using aluminum chloride hexahydrate (AlCl3  6H2O) dissolved in a cetyltrimethylammonium (CTABr) and sodium
hydroxide (NaOH) under intense agitation. Next, tetraethylorthosilicate (TEOS) was slowly added. The nal product was then dried
in an oven at 105 C for 24 h to give catalysts Al-MCM-41 with
Si/Al ratios of 8, 16, and 32, designated Al-MCM-41(8),
Al-MCM-41(16) Al-MCM-41(32), respectively.
Samples Al-MCM-41(8)S, Al-MCM-41(16)S and Al-MCM41(32)S were submitted to thermal treatment for 7 h in a tubular
oven at 550 C at heating rate of 15 C min1, 1 h under a continuous ow of nitrogen and 6 h under a continuous ow of compressed air at a gas ow rate of 200 mL min1 to eliminate the
surfactant residue (CTABr) from the pores of the aluminosilicate.
After the thermal treatment, the samples were designated AlMCM-41(8)C, Al-MCM-41(16)C, and Al-MCM-41(32)C. For the
catalysis tests, catalyst ionic exchange was carried out in NH4NO3
0.5 M at 80 C for 20 h. Next, the catalysts were calcined in an oven
at 480 C for 3 h a static ow air, giving the protonated solid.
2.2. Characterization
Atomic absorption spectroscopy analyses were performed using
a VARIAN model SPECTRA 55 (k = 309.3 nm, slit u = 0.5 nm). TG and
DTA curves were obtained in static air atmosphere over the temperature range 25850 C at 10 C min1 in a THERMAL SCIENCE ther-

Table 1
Quantity of Al in the solid Al-MCM-41.
Samples

Al-MCM-41 (8)
Al-MCM-41 (16)
Al-MCM-41 (32)

Weight of Al (mg)
Theoretical

Obtained

156.8
78.4
39.2

157.2
77.0
38.5

463

A.C. Carmo et al. / Fuel 88 (2009) 461468

crystallographic patterns based on small reection angles from


the (1 0 0), (1 1 0), and (2 0 0) crystal planes, characteristic of the
mesoporous solid aluminosilicate Al-MCM-41. The crystallographic data for Al-MCM-41 are given in Table 2 for the different
Si/Al ratios.
An increase in the amount of aluminum incorporated into the
mesopores leads to disorder in the structural arrangement of AlMCM-41. This leads to broadening of the (1 0 0), (1 1 0), and (2 0 0)
peaks. We observe that the hexagonal orientation of Al-MCM-41
is drastically affected by the amount of aluminum present, producing alterations in the porous structure of the solid, although without changing its primary structure [24]. The surface area values,
and pore diameter and volume are given in Table 2. The N2 adsorption and desorption isotherms are shown in Fig. 3.
We note an agreement between the results obtained for AlMCM-41 in the present work and literature data [25]. The AlMCM-41 aluminosilicates synthesized in this work show high
values for their surface area and pore volume, with average pore
diameter in the mesopores range. The uniformity of the pore lattice
is evaluated according to the N2 adsorption/desorption isotherm
behavior. The materials display a structure that corresponds to
Al-MCM-41. It was also observed that all catalysts showed a
type-IV isotherm behavior, which is characterized by the appearance of three adsorption stages, namely the formation of the
monolayer by pore wall coating (0.30 < p/p0 < 0.40), pore lling
by capillary condensation, which starts at the adsorption multilayer characterized by a signicant increase in relative pressure
(p/p0 < 0.30) and the coating of the solid by the formation of the
multilayers in regions outside the pore (p/p0 > 0.4) [22,25,26]. In
this analysis, we also observed the inuence of the increase in
the amount of aluminum on the pore structure with a gradual decrease in diameter and pore volume.

100
6

Al-MCM-41 (8)
Al-MCM-41 (16)
Al-MCM-41 (32)

90

70

60

dT (C)

Weight (%)

80

50
-2
40
200

400

600

800

Temperature (C)
Fig. 1. TG/DTA curves of Al-MCM-41.

(or combustion) of organic components due to the removal of


surfactant molecules (CTABr) from the material pores. The third
mass loss (9.0%) occurs between 400 and 850 C. The mass loss in
this region is related to water from the condensation of the silanol
groups to form siloxane groups and is represented by the exothermic peak close to 550 C [22]. Several studies on the removal of
the micellar nucleus show that the degradation of the surfactant
at high temperatures occurs by a systematic mechanism with its
most advanced stage in the temperature interval from 200 to
300 C [23].
The diffraction patterns of the synthesized (S) and calcined (C)
catalyst Al-MCM-41 forms are given in Fig. 2. The samples present

Intensity (a.u.)

(100)

(100)

(110) (200)
Al-MCM-41 (8)S

Al-MCM-41 (16)S

Intensity (a.u.)

Al-MCM-41 (32)S

(110)

(200)

Al-MCM-41 (8)C

Al-MCM-41 (16)C

Al-MCM-41 (32)C

5
2

Fig. 2. X-ray diffraction pattern of Al-MCM-41 samples.

Table 2
Physical properties of catalyst Al-MCM-41 with 8, 16, and 32 Si/Al.
Samples

2h

d100 ()

a0 ()

Surface area (m2/g)

APD ()

Pore volume (cm3/g)

Al-MCM-41 (8)
Al-MCM-41 (16)
Al-MCM-41 (32)

2.50
2.42
2.21

33
36
40

38
42
46

713
792
815

34
36
36

0.67
0.71
0.89

464

A.C. Carmo et al. / Fuel 88 (2009) 461468

The vibrational spectra of the solid Al-MCM-41 with Si/Al


ratios of 8, 16, and 32 were obtained by observation of their
characteristic absorption bands and changes caused by the thermal
treatment used in the production of the nal product.
The spectra obtained before calcination show the presence of
stretching absorption bands at 29192922 cm1, 28512853 cm1
and 14711476 cm1 (symmetric and asymmetric). These relate
to the angular deformations in the CH bond plane of the CH2 and

CH3 groups of the molecules of the organic templates. It was also


observed that before calcination the catalysts displayed absorption
bands relating to SiO bond stretching centered at 1217
1223 cm1, 10331043 cm1 and 685790 cm1, which remain,
with minor modication, after the thermal treatment.
According to Selvaraj et al. [27], the efciency of the calcination
process may be inferred by the behavior of the spectra upon calcination. It may be observed that there is an absence of bands from

650
Si/Al=8
Si/Al=16
Si/Al=32

600
550
500

Volume (cm /g)

450
400
350

120
110
100

200

3
-1

80

-1

250

90
70

Pore volume (cm .A .g .10 )

300

60

150

50
40
30
20
10
0
0

20

40

60
80
100
120
Pore Diameter (A)

140

160

180

100
0.0

0.2

0.4

0.6

0.8

1.0

Relative Pressure (p/p0)


Fig. 3. Adsorption/desorption isotherms of Al-MCM-41 and its pore distribution.

Fig. 4. SEM micrographs of Al-MCM-41 catalysts.

1.2

465

A.C. Carmo et al. / Fuel 88 (2009) 461468

(CH3) and (CH2) groups relative to the surfactant chain. It is important to point out that, even after calcination, the primary structure
of Al-MCM-41 undergoes only small alterations, which is evidenced by the modest changes in the spectra. The appearance of
the bands between 500 and 1200 cm1 is attributed to structural
vibrations and the signals between 1088 and 1230 cm1 from the
TOT bonds (T = Si or Al). Normally these signals are attributed
to SiOH vibrations. However, when several metals are incorporated, the intensity of this band tends to increase, thus demonstrating good incorporation of the metals into the mesopore structure.
The morphology of the Al-MCM-41 catalyst particles before and
after calcination was studied using the SEM micrographs presented
in Fig. 4.
It can be observed in the images that the macrostructure of catalysts Al-MCM-41 (8) and Al-MCM-41 (16) remains intact even
after the calcination process, thus conrming the high thermal stability of Al-MCM-41. We also observe the presence of particles with
varied morphology (agglomerated and spheroid), apparently homogeneous, thus providing good evidence for incorporation of aluminum to the structure. The same conclusion holds for the images of
catalyst Al-MCM-41 (32), whose particles also have good size and
well-dened morphology, though without spherical particles. Analysis of particle size distribution gave 116 lm for Al-MCM-41 (8),

280 lm for Al-MCM-41 (16), and 1140 lm for Al-MCM-41 (32).


According to Ekloff et al. [28], this difference in products obtained
by the same synthesis method is because the precipitation of
MCM-41 is not uniform due to variations of the reaction medium
pH, which may delay the formation of the structure.
TEM micrographs of calcined Al-MCM-41 catalyst are given in
Fig. 5.
The images of calcined Al-MCM-41 present a well-dened hexagonal arrangement with a fairly uniform pore structure. We can
also see the inuence on the structure of the amount of aluminum,
whose decrease results in less ordered materials. However, the
crystalline structure does not occur throughout the solid structure,
as disordered regions are also present. This behavior, the presence
of disordered regions, lamellar phases, and ngerprint-like structures, has been reported in the literature [29]. Kang et al. [30] obtained images of Al-MCM-41 with metakaolinite as a source of
aluminum. They show the formation of a honeycomb-like structure from parallel channels with a highly ordered porous lattice,
complementing XRD and N2 adsorption analysis results.
3.2. Catalytic activity
The catalytic activity of Al-MCM-41 for the esterication of palmitic acid with methanol, ethanol and isopropanol was determined
from the acidity index of the nal product. Fig. 6 shows the effect
of each catalyst on the esterication process.
It can be veried that the esterication of palmitic acid follows
this order of effectiveness: methanol > ethanol > isopropanol. This
was expected as the carbon chains of each alcohol have different
reactivities. Kirumakki et al. [31] report that the size and type of
the alcohol chain inuence the esterication with zeolites Hb,
HY, and HZSM-5. We also clearly note the inuence of the Si/Al ratio on the reaction with both alcohols, where the largest conversion values are obtained with low Si/Al ratios due to the
Brnsted acid sites having been incorporated into the structure
of MCM-41 by the introduction of aluminum [32]. Table 3 presents
some esterication results found in the literature. We can see the
high activity of Al-MCM-41 in the esterication processes, mainly
in the gas phase, in which the reaction occurs mainly in the catalyst pores. For the esterication of palmitic acid aimed at the production of biodiesel, Al-MCM-41 is active for low Si/Al ratios, that
is, for large concentrations of catalytic sites.

80

Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

70

Conversion (%)

60
50
40
30
20
10
0

Methanol

Fig. 5. TEM micrographs of Al-MCM-41.

Ethanol

Isopropanol

Fig. 6. Effect of Al-MCM-41 catalysts on the production of biodiesel by


esterication.

466

A.C. Carmo et al. / Fuel 88 (2009) 461468

Table 3
Esterication reaction with different acidic catalysts.
Acid

Alcohol

Alcohol/acid

Temperature (C)

Catalyst

Acetic
Acetic
Acetic
Palmitic
Palmitic
Palmitic

Amilic
n-Butanol
n-Propanol
Methanol
Ethanol
Isopropanol

1/2
60/1
60/1
60/1

250
125
150
130
130
130

Al-MCM-41
Al-MCM-41
Al-MCM-41
Al-MCM-41
Al-MCM-41
Al-MCM-41

(Si/Al = 100)
(Si/Al = 25)
(Si/Al = 30)
(Si/Al = 8)
(Si/Al = 8)
(Si/Al = 8)

Conversion (%)

Refs.

91.0
87.3
83.7
79.0
67.0
59.0

[36]
[32]
[21]
(*)
(*)
(*)

Present work.

3.3. Inuence of reaction time


The esterication conditions used with palmitic acid and the
different alcohols were 130 C, alcohol/acid molar ratio of 60,
0.6 wt% catalyst (total mixture). Under these conditions we
determined the degree of conversion as a function of time. Fig. 7
shows the conversion proles relative to the reaction time for
the different alcohols. In the reaction with methanol, the conversion is about 79% for Si/Al = 8, falling to 67% for Si/Al = 16, and to
56% for Si/Al = 32. For ethanol, the conversion is 67% for Si/Al = 8,
57% for Si/Al = 16, and 45% for Si/Al = 32. The trend holds for
isopropanol with 59% for Si/Al = 8, 48% for Si/Al = 16, and 36% for
Si/Al = 32. Also illustrated in Fig. 7, the time period of 1 h is

sufcient for the maximum conversion to be reached independently of the alcohol that was used.
The order of the esterication reaction may be determined following the classical denitions of chemical kinetics. Considering
the conditions employed in the process, palmitic acid is a limiting
reagent. Its transformation into esters is followed by the acidity index. It is proposed that the acid is completely converted so allowing a denition of the order of the reaction by tting to elementary
kinetic theory. Fig. 8 gives possible ts of the esterication reaction
of palmitic acid assuming rst-order kinetics.
The tting reveals a linear relation between all experimental
data when (-ln (1-conversion)) is plotted as a function of reaction
time, thus establishing the existence of a rst-order dependence

b
80

80

a
60

40

Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

40

Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

20

20

0
0

30

60

90

120

30

60

90

Time (min)

Time (min)

c
60
50

Conversion (%)

Conversion (%)

60

40
30
20
10

Si/Al = 8
Si/Al = 16
Si/Al = 32
Without catalyst

0
0

30

60

90

120

Time (min)
Fig. 7. Conversion proles of the esterication of palmitic acid with (a) methanol, (b) ethanol and (c) isopropanol.

120

467

A.C. Carmo et al. / Fuel 88 (2009) 461468

b 1.0
Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

1.2

1.0

0.8

0.8

-ln(1-X)

Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

0.6

0.6
0.4

0.4
0.2

0.2

0.0

0.0

10

15

20

25

30

10

Time(min)

15

20

25

30

Time(min)

0.8

Si/Al=8
Si/Al=16
Si/Al=32
Without catalyst

0.7
0.6

-ln(1 - X)

0.5
0.4
0.3
0.2
0.1
0.0
0

10

15

20

25

30

Time(min)
Fig. 8. Esterication of palmitic acid with (a) methanol, (b) ethanol and (c) isopropanol. Reaction tting to rst-order kinetics.

3.4. Low quality oil catalyst test


The reaction performance of the Al-MCM-41 with Si/Al = 8 was
evaluated for the esterication of free fatty acids in a fatty acid/triglyceride mixture. A mixture of 17 wt% oleic acid in palm oil was
used as the model high free fatty acid feed. In Fig. 9 it can observed
a maximum conversion of 40% at a temperature of 130 C and in a
time period of 1 h. The percentage of conversion can be raised by
manipulating such variables as time and temperature. For example,
at a temperature of 180 C for a time period of 2 h, the removal of FFA
observed is 60%. These values can be considered high when com-

65
60
55

1h
2h

50
Conversion (%)

between the reaction rate and the concentration of carboxylic acid


for the esterication of palmitic acid. The regression coefcients of
the straight lines show good ts to rst-order kinetics.
Several reports in the literature establish rst-order kinetics for
this esterication reaction [3133]. In contrast, Tanaka et al. [34]
demonstrated second-order kinetics for the esterication of ethanol
in the presence of zeolites. This difference in reaction order, mainly
for heterogeneous reactions, has often been attributed to the
adsorption mechanism prevailing at the reagentcatalyst interface.

45
40
35
30
25
20
15
10
5
0
130C

180C
Temperature (C)

Fig. 9. Relation of conversion of FFA and reaction temperature: reaction time 1 h


and 2 h.

468

A.C. Carmo et al. / Fuel 88 (2009) 461468

pared with other catalysts that obtain greater yield, but with sites of
stronger acids. For example, as presented by Mbaraka et al. [4],
organosulfonic acid-functionalized mesoporous silicas has reached
a percentage of conversion of fatty acids to methyl esters of approximately 80%. This percentage approximates the one observed by
Chung et al. [35] with zeolite catalysts. Both organosulfonic acidfunctionalized mesoporous silicas and zeolite catalysts are stronger
acids sites than MCM, however, this one emerges as a more simple
alternative for the pretreatment process of low quality oils.
4. Conclusions
The solid catalyst Al-MCM-41 was obtained with different Si/Al
ratios by synthesis at room temperature and characterized in terms
of its thermal, structural, and morphological aspects. These materials are active in the esterication of palmitic acid. The catalyst conversion rates of Al-MCM-41 with Si/Al = 8 were 79%, 67%, and 59%
for methanol, ethanol, and isopropanol, respectively. The esterication reaction of palmitic acid in a xed atmosphere allows observation of the inuence of the amount of incorporated metal and
consequently the concentration of acidic sites in each catalyst.
The reaction conditions, particularly agitation and temperature,
play a key role in the determination of the on reaction kinetics;
effective and uniform agitation results in good miscibility of reagents and thus favors their diffusion to catalytic sites. At the temperature adopted, 130 C, the negative effect of the water formed in
the reaction barely interfered with the nal degree of conversion.
The experimental data obtained showed a good t to pseudo-rst
order kinetics. However, further studies of the combined effect of
other variables that certainly inuence the process will be required
to conrm not only the reaction order, but also other kinetic parameters relevant for the reaction such as reaction rate, rate constant
and activation energy. The data obtained also present Al-MCM-41
as an alternative for removal of free fatty acids from low quality oils.
Acknowledgement
The nancial support of Laboratrio de Pesquisa e Anlise de
Combustveis (LAPAC/UFPa) granted to this project.
References
[1] Liu Y, Lotero E, Goodwin Jr JG. Effect of carbon chain length on esterication of
carboxylic acids with methanol using acid catalysis. J Catal 2006;243:2218.
[2] Yadav GD, Thathagar MB. Esterication of maleic acid with ethanol over cation
exchange resin catalysts. React Funct Polym 2002;52:99110.
[3] Marchetti JM, Miguel VU, Errazu AF. Heterogeneous esterication of oil with
high amount of free fatty acids. Fuel 2007;86:90610.
[4] Mbaraka IK, Radu DR, Lin VS-Y, Shanks BH. Organosulfonic acid-functionalized
mesoporous silicas for the esterication of fatty acid. J Catal 2003;219:32936.
[5] Marchetti JM, Errazu AF. Esterication of free fatty acids using sulfuric acid as
catalyst in the presence of triglycerides. Biomass Bioenerg 2008;32:8925.
[6] Tanagaki A, Toda M, Okamura M, Kondo JN, Hayashi S, Domen K, et al.
Esterication of higher fatty acids by a novel strong solid acid. Catal Today
2006;116:15761.
[7] Lpez DE, Suwannakarn K, Bruce DA, Goodwin Jr JG. Esterication and
transesterication on tungstated zirconia: effect of calcination temperature. J
Catal 2007;247:4350.
[8] Biz S, White MG. Effect of post-synthesis hydrothermal treatments on the
adsorptive volume of surfactant-templated mesostructures. Micropor Mesopor
Mater 2000;40:15971.
[9] Ni J, Meunier FC. Esterication of free fatty acids in sunower oil over solid
acid catalysts using batch and xed bed-reactors. Appl Catal A: Gen
2007;333:12230.

[10] Ramu S, Lingaiah N, Prabhavathi Devi BLA, Prasad RBN, Suryanarayana I, Sai
Prasad PS. Esterication of palmitic acid with methanol over tungsten oxide
supported on zirconia solid acid catalysts: effect of method of preparation of
the catalyst on its structural stability and reactivity. Appl Catal A: Gen
2004;276:1638.
[11] Beck JS, Vartuli JC, Roth WJ, Leonowicz ME, Kresge CT, Schmitt KD, et al. A new
family of mesoporous molecular sieves prepared with liquid crystal templates.
J Am Chem Soc 1992;114:1083443.
[12] Twaiq F, Zabidi NAM, Mohamed AR, Bhatia S. Catalytic conversion of palm oil
over mesoporous aluminosilicate MCM-41 for the production of liquid
hydrocarbon fuels. Fuel Process Technol 2003;84:10520.
[13] Selvaraj M, Jeon SH, Han J, Sinha PK, Lee TG. A novel route to produce 4-tbutyltoluene by t-butylation of toluene with t-butylalcohol over mesoporous
Al-MCM-41 molecular sieves. Appl Catal A: Gen 2005;286:4451.
[14] Climent MJ, Corma A, Iborra S, Miguel S, Primo J, Rey F. Mesoporous materials
as catalysts for the production of chemicals: synthesis of alkyl glucosides on
MCM-41. J Catal 1999;183:7682.
[15] Barrault J, Pouilloux Y, Clacens JM, Vanhove C, Bancquart S. Catalysis and ne
chemistry. Catal Today 2002;75:17781.
[16] Wloch J, Rozwadowski M, Lezanska M, Erdmann K. Analysis of the pore
structure of the MCM-41 materials. Appl Surf Sci 2002;191:36874.
[17] Bhagiyalakshmi M, Shanmugapriya K, Palanichamy M, Arabindoo B,
Murugesan V. Esterication of maleic anhydride with methanol over solid
acid catalysts: a novel route for the production of heteroesters. Appl Catal A:
Gen 2004;267:7786.
[18] Shylesh S, Samuel PP, Singh AP. Synthesis of hydrothermally stable
aluminium-containing ethane-silica hybrid mesoporous materials using
different aluminium sources. Micropor Mesopor Mater 2007;100:2508.
[19] Ajaikumar S, Pandurangan A. Esterication of alkyl acids with alkanols over
MCM-41 molecular sieves: inuence of hydrophobic surface on condensation
reaction. J Mol Catal A: Chem 2007;266:110.
[20] Luque R, Campelo JM, Luna D, Marinas JM, Romero AA. Catalytic performance
of Al-MCM-41 materials in the N-alkylation of aniline. J Mol Catal A: Chem
2007;269:1906.
[21] Gokulakrishnan N, Pandurangan A, Sinha PK. Esterication of acetic acid with
propanol isomers under autogeneous pressure: a catalytic activity study of AlMCM-41 molecular sieves. J Mol Catal A: Chem 2007;263:5561.
[22] Jermy BR, Pandurangan A. Al-MCM-41 as an efcient heterogeneous catalyst in
the acetalization of cyclohexanone with methanol, ethylene glycol and
pentaerythritol. J Mol Catal A: Chem 2006;256:18492.
[23] Ryczkowski J, Goworek J, Gac W, Pasieczna S, Borowiecki T. Temperature
removal of templating agent from MCM-41 silica materials. Thermochim Acta
2005;434:28.
[24] Mokaya R, Jones W. Physicochemical characterisation and catalytic activity of
primary amine templated aluminosilicate mesoporous catalysts. J Catal
1997;172:21121.
[25] Souza MJB, Arajo AS, Pedrosa AMG, Marinkovic BA, Jardim PM, Morgado Jr E.
Textural features of highly ordered Al-MCM-41 molecular sieve studied by Xray diffraction, nitrogen adsorption and transmission electron microscopy.
Mater Lett 2006;60:26825.
[26] Leofanti G, Padovan M, Tozzola G, Venturelli B. Surface area and pore texture of
catalysts. Catal Today 1998;41:20719.
[27] Selvaraj M, Pandurangan A, Seshadri KS, Sinha PK, Lal KB. Synthesis,
characterization and catalytic application od MCM-41 mesoporous
molecular sieves containing Zn and Al. Appl Catal A: Gen 2003;242:34764.
[28] Ekloff GS, Rathousky J, Zukal A. Controlling of morphology and
characterization of pore structure of ordered mesoporous silicas. Micropor
Mesopor Mater 1999;27:27385.
[29] Ciesla U, Schth F. Ordered mesoporous materials. Micropor Mesopor Mater
1999;27:13149.
[30] Kang F, Wang Q, Xiang S. Synthesis of mesoporous Al-MCM-41 materials using
metakaolin as aluminum source. Mater Lett 2005;59:14269.
[31] Kirumakki SR, Nagaraju N, Chary KVR. Esterication of alcohols with acetic
acid over zeolites H, HY and HZSM5. Appl Catal A: Gen 2006;299:18592.
[32] Jermy BR, Pandurangan A. Catalytic application of Al-MCM-41 in the
esterication of acetic acid with various alcohols. Appl Catal A: Gen
2005;288:2533.
[33] Palani A, Pandurangan A. Esterication of terephthalic with methanol
over mesoporous Al-MCM-41 molecular sieves. J Mol Cat A: Chem 2006;
245:1015.
[34] Tanaka K, Yoshikawa R, Ying C, Kita H, Okamoto KI. Application of zeolite
membranes to esterication reactions. Catal Today 2001;67:1215.
[35] Chung K-H, Chang D-R, Park B-G. Removal of free fatty acid in waste frying oil
by esterication with methanol on zeolite catalysts. Bioresource Technol
2008;99:743843.
[36] Palani A, Pandurangan A. Esterication of acetic acid over mesoporous
Al-MCM-41 molecular sieves. J Mol Catal A: Chem 2005;226:12934.

Vous aimerez peut-être aussi