Você está na página 1de 11

Applied Catalysis A: General 462463 (2013) 196206

Contents lists available at SciVerse ScienceDirect


Applied Catalysis A: General
j our nal homepage: www. el sevi er . com/ l ocat e/ apcat a
Mesoporous nanocrystalline sulfated zirconia synthesis and its
application for FFA esterication in oils
Vishwanath Ganpat Deshmane
a
, Yusuf Gbadebo Adewuyi
b,
a
Mechanical Engineering Department, North Carolina Agricultural and Technical State University, Greensboro, NC 27411, USA
b
Chemical, Biological and Bioengineering Department, North Carolina Agricultural and Technical State University, Greensboro, NC 27411, USA
a r t i c l e i n f o
Article history:
Received 20 December 2012
Received in revised form3 May 2013
Accepted 5 May 2013
Available online xxx
Keywords:
Sulfated zirconia
Esterication
Solid acid catalyst
Free fatty acids
Biodiesel
Catalyst characterization
a b s t r a c t
Mesoporous nanocrystalline sulfated zirconia catalyst has been prepared from zirconium hydroxide syn-
thesized at different digestion/hydrothermal treatment times. Sulfuric acid and chlorosulfonic acid were
usedas two different sulfonating agents. The effect of digestion time, sulfonating agent andthe calcination
temperature on structural, textural and catalytic properties of the prepared catalyst were investigated
in details using nitrogen adsorptiondesorption (BET), ammonia temperature programmed desorption
(NH
3
-TPD), X-ray diffraction (XRD), thermogravimetry and differential scanning calorimetry (TGADSC),
and Fourier transform infrared spectroscopy (FTIR). The sulfated zirconia prepared at various digestion
times and two different sulfonating agents were tested for the esterication of free fatty acid (FFAs) in
soybean oil (prepared by mixing oleic acid in soybean oil) as model reaction. Sulfated zirconia catalyst
prepared with 3 h digestion time and 600

C calcination temperature using chlorosulfonic acid showed


the highest catalytic activity with about 85% conversion in just 80min of esterication time at 60

C
reaction temperature.
2013 Elsevier B.V. All rights reserved.
1. Introduction
Biodiesel also called fatty acid methyl ester is a clean-burning,
renewable fuel produced from vegetable oils, animal fats and
recycledcookingoil andgreases, etc. It is not onlybiodegradablebut
also free of sulfur, making it a cleaner burning fuel than petroleum
diesel withreducedemissionof SOx, CO, unburnt hydrocarbons and
particulate matter [1]. Excellent lubricating properties that extend
engine life, superior cetane number, ash point compared to con-
ventional diesel and acceptable cold lter plugging point (CFPP) are
some of the attributes that make biodiesel very attractive alterna-
tive fuel [2,3].
The major hurdle in the use of biodiesel for replacing conven-
tional petroleumfuels is its higher cost. The two main factors that
affect the cost of biodiesel are the cost of raw materials and the
processingcost suchas catalysts andequipments [4]. Therawmate-
rials account for over 6075% of the biodiesel production expenses.
The potential solution to this problemis the utilization of lowvalue
alternative feedstocks of varying type, quality and cost. For exam-
ple, the cost of waste cooking oil is 23 times lower than virgin
oils. Thus the utilization of less expensive feedstocks such as ani-
mal fat, waste cooking oil, yellowand brown grease is expected to

Corresponding author. Tel.: +1 336 334 7564x107; fax: +1 336 334 7417.
E-mail address: adewuyi@ncat.edu (Y.G. Adewuyi).
appreciably reduce the biodiesel cost [5]. However, many of these
alternative feedstocks may contain high levels of free fatty acids
(FFA), water, or insoluble matter, which affect biodiesel production
[6]. Synthesis of biodiesel via transestericationreactionwithfeed-
stocks having higher FFA and moisture is complicated. During the
reaction, the feedstocks undergo saponication reaction leading to
soaps formation resulting in reduced biodiesel yields, especially
when alkaline catalysts are used. Furthermore, the soap formation
also leads to the catalyst consumption, lowering catalytic efciency
and increase in the viscosity of reaction mixture and gel formation
requiring additional purication steps [5]. These problems could
potentially be eliminated via the use of heterogeneous acid cata-
lysts due to their lower susceptibility to FFAs and moisture content
in the oil [7]. Also, catalysts can be easily separated fromthe reac-
tion products with much more simplied product separation steps
resulting in high yields of methyl esters and decrease of catalyst
cost due to the possibility of catalyst regeneration.
To date, several solid acid catalysts have been reported for
biodiesel synthesis, including zeolites (e.g. H-ZSM-5, Y and Beta),
ion exchange resins (e.g. Amberlist 12, a styrene based sul-
fonic acid and Naon-NR-50, a copolymer of tetrauoroethene
and peruoro-2-(uorosulphonyle-thoxy) propyl vinyl ether) and
metal oxides modied with sulfate ions (SO
4
2
/M
x
O
y
, such as
SO
4
2
/ZrO
2
, SO
4
2
/SnO
2
, SO
4
2
/TiO
2
, SO
4
2
/WO
3
), etc. [8,9]. Zeo-
lite catalysts with small (micron-sized) pores are not suitable for
biodiesel manufacture because of the diffusion limitations induced
0926-860X/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2013.05.005
V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206 197
by the large fatty acid molecules. Ion-exchange resins are active
strong acids, but have a low thermal stability which is problem-
atic as the esterication reaction might require high temperatures.
Metal oxides modied with sulfate ions and especially sulfated
zirconia, has high boiling point, strength, toughness, good cor-
rosion resistant in acidic and alkaline environment. In addition
sulfated zirconia has very high activity, selectivity and stability
making it a promising candidate not only for the estericationreac-
tion but also for number of industrially important reactions such
as hydrocarbon isomerization, methanol conversion to hydrocar-
bons, alkylation, acylation, etherication, condensation, nitration,
cyclization and FisherTropsch synthesis [1012]. However, the
catalytic and structural properties of the sulfated zirconia depends
on number of factors including Zr(OH)
2
preparation method, pre-
cursor used, precipitation pH, precursor concentration, type of
sulfonating agent, catalyst pretreatment, and calcination [1315].
In an earlier work, nanocrystalline mesoporous Zr(OH)
2
pow-
der with very high surface area was synthesized using ethylene
diamine and zirconyl chloride octahydrate. Ethylene diamine used
as precipitating agent also acted as a colloidal protecting agent.
The results of the effects of various process parameters such as
precipitation pH, precursor concentration, time of hydrothermal
treatment and calcination temperature on the structural and tex-
tual properties of zirconiumoxide were discussed in detail [16].
Yadav and Murkute [17] previously reported the use of chloro-
sulfonic for the sulfonation of zirconia powder. It was showed that
the chorosulfonic acid treated zirconia possesses more sulfate ions,
stability and activity compared to the sulfated zirconia synthesized
using sulfuric acid. In this study we report the synthesis of meso-
porous nanocrystalline sulfated zirconia with high surface area
and acidity. Mesoporous zirconium hydroxide prepared at differ-
ent digestion times of 0, 1, 3, 6, 12, 24 and 48h were sulfonated
by wet impregnation method using sulfuric acid and chlorosul-
fonic acid as two different sulfonating agents. To the best of our
knowledge, nostudies have beenreportedinthe literature showing
the effect of digestion/hydrothermal treatment on the sulfonation
process and acidity of nal sulfated zirconia catalyst. The effect of
preparation conditions such as digestion time, sulfonating agent
and the calcination temperature on the structural phases, textu-
ral characteristics and the number and types of available active
acidic sites on the surface of the nal sulfated zirconia were inves-
tigated using nitrogen adsorptiondesorption (BET), ammonia
temperature programmed desorption (NH
3
-TPD), X-ray diffraction
(XRD), thermogravimetry and differential scanning calorimetry
(TGADSC), and Fourier transform infrared spectroscopy (FTIR).
The sulfated zirconia prepared at the best synthesis conditions was
tested for the esterication of free fatty acid in soybean oil (pre-
pared by mixing oleic acid in soybean oil) as model reaction.
2. Experimental
2.1. Chemicals
Zirconyl chloride octahydrate (ZrOCl
2
8H
2
O), 98+% pure,
ethylenediamine (H
2
NCH
2
CH
2
NH
2
), 99%, extra pure, sulfuric acid,
97%, ethylene dichloride were purchased fromAcros Organics, NJ,
USA. Chlorosulfonic acid, 99% was obtained fromAlfa Aesar, Ward
Hill, USA. The water used at all stages of the experiments was puri-
edusing a Mill-QAdvantage A10 withElix 5 systemobtainedfrom
Millipore Corporation (Bedford, MA, USA).
2.2. Catalyst synthesis
The method for the synthesis of zirconium hydroxide (8090%
yield) has been discussed in details previously [16]. Zirconyl
chloride octahydrate and ethylene diamine are used as zirco-
nium precursor and precipitating agent, respectively, as reported
by Dsouza et al. [18]. Sulfated zirconia was prepared from this
material using two different methods based upon two different
sulfonating agents, i.e., sulfuric acid and chlorosulfonic acid. In the
case of sulfuric acid, 1g of the dried again as prepared zirconium
hydroxide power was mixed with 15ml of 1N H
2
SO
4
and then
stirredwithmagnetic stirrer for about 10min followedbythe ltra-
tion and air drying. The air dried material was then dried in oven
for 24h at 110

C. In the case of chlorosulfonic acid, 1g of dried


as prepared zirconium hydroxide power was immersed in 15ml,
0.5M solution of chlorosulfonic acid in ethylene dichloride. After
about 30min, ethylene dichloride was evaporated in an oven at
80

Cfor 20handthendriedcompletely for 24hat 110

C. The sam-
ples prepared by using both methods were then calcined at 600

C
and 650

C temperature for 2h with controlled heating and cool-


ing rates of 0.5

C/min and 1

C/min respectively, in the presence


of air. The two different samples were denoted as SZ and CSZ for
sulfated zirconia prepared using sulfuric acid and chlorosulfonic
acid, respectively. Finally, the prepared samples were character-
ized by using different analytical and instrumentation techniques
described in the following section.
2.3. Catalyst characterization
The BET surface area, total pore volume and pore size dis-
tribution of the catalyst were determined with AUTOSORB-1C,
ChemisorptionPhysisorption analyzer (Quantachrome Instru-
ments, Boynton Beach, FL, USA). Surface area was calculated by
using BET equation from the adsorption branch of the isotherm
in a relative pressure range of 0.070.3. The pore size dis-
tribution was calculated from desorption branches using the
BarrettJoynerHalenda (BJH) method [19]. The total pore volume
was derived based on the amount of N
2
adsorbed at a relative pres-
sure close to unity. The acidity of the sulfated zirconia catalyst was
measuredby using ammonia temperature programmeddesorption
(NH
3
-TPD) experiments. Thermal conductivity detector (TCD) con-
nected to the AUTOSORB-1C was used to measure the ammonia
desorption prole. In a typical run, 0.25g of the catalyst (sand-
wiched between two small wads of glass wool) was placed into the
chemisorptions cell and heated to 140

C at 20

C/min under the


heliumowfor 30min to remove adsorbed components. Later the
samplewas cooledto100

Candsaturatedwithammoniabyexpos-
ing the sample to 100% NH
3
for 10min. Physisorbed ammonia was
removed by purging the sample withheliumgas for 30min. Finally,
the temperature was ramped to 600

C at a rate of 20

C/min and
evolvedammonia was quantiedby thermal conductivity detector.
Thermo-gravimetric (TGA) and differential scanning calorime-
try analysis (DSC) were carried out using a SDT Q600 V20.4 Build
14 system(TA Instruments, NewCastle, DE, USA). The heating was
carried out in an air environment. The air owrate was maintained
at 100ml/min and the heating rate was 10

C/min.
Infrared absorptiontransmission spectra were obtained using
an FTIR spectrometer (TENSOR 27, Bruker Optics, Inc., Billerica,
MA) with HeNe laser source and a room-temperature deuterated l-
alanine triglycine sulfate detector (DLATGS detector). The FTIR was
equipped with an ATR sampling accessory, MIRacle ATR set with
Diamond crystal assembly fromPIKE Technologies (PIKE Technolo-
gies, Inc., Madison, WI). All spectra were collected at 201

C using
an average of 16 scans and with a spectral resolution of 2cm
1
.
The background spectra were obtained using a clean ATR accessory
with a continuous dry, CO
2
-free air purge from a laboratory gen-
erator (Parker-Balston, Haverhill, MA) to remove moisture. During
the analysis, a small amount of powderedsample was placedonthe
crystal and pressured with high pressure clamp to get the intimate
contact between the sample and crystal surface.
198 V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206
0 200 400 600 800 1000
75
80
85
90
95
100 0 h
1 h
3 h
6 h
12 h
24 h
48 h
Temperature (C)
W
e
i
g
h
t

(
%
)
-2.0
-1.5
-1.0
-0.5
0.0
0.5
H
e
a
t

F
l
o
w

(
W
/
g
)
Fig. 1. TGADSC plot for the zirconium hydroxide prepared at different digestion
times.
The XRDpatterns were recorded on D8 DISCOVER X-ray diffrac-
tometer fromBruker (Bruker Optics, Inc., Billerica, MA) with a PSD
detector using Cu K radiation generated at 40mA and 40kV at
the scanning rate of 0.01

/s. The crystal sizes of the samples were


determined by using Scherrer equation [20] given by the following
equation:
=
0.9
cos
(1)
where is the crystal size, is the wavelength of the Cu K radia-
tion, is the full width half maximumof the respective peak and
is the Braggs angle of diffraction.
2.4. Catalyst testing
The catalytic activity/performance of the prepared sulfated zir-
conia catalyst was determined by testing it for the esterication of
oleic acid in soybean oil (10% oleic acid) as model reactant for used
oil hereafter mentioned as acid oil. The reaction was carried out in
a cylindrical jacketed glass reactor with four necks. It was equipped
with condenser to reux the methanol evaporated during the reac-
tion. A thermocouple with digital temperature indicator was used
to measure the temperature of the reaction mixture. The stirring
was carried out using a magnetic stirrer (396W, StableTemp, Cole
Parmer) at 1000rpm. To start the experiments, 42ml acid oil and
18ml methanol (methanol/oil: 9/1) were added to the reactor. The
mixture was heated to 60

C by circulating the hot water through


a reactor jacket with continuous stirring. Then 2% (wt % of acid oil)
of catalyst was added to start the reaction. Samples (2ml aliquot)
were removed fromthe reaction mixture at specied times during
the progress of reaction and immediately cooled to 1012

C tem-
perature. The cooled sample was then centrifuged to separate the
solid catalyst fromthe liquid reaction mixture. Approximately, 1g
of the separated liquid phase was dissolved in 5ml of 2-propanol
to make a homogeneous solution which was then titrated against
0.05N KOH solution in the presence of phenolphthalein indicator
to determine the acid value (mg of KOH required to neutralize 1g
of the sample).
3. Results and discussion
3.1. TGADSC
The TGADSC plots for the zirconiumhydroxide prepared at dif-
ferent digestion times (048h) are shown in Fig. 1. Two weight
0 200 400 600 800 1000
80
85
90
95
100
105
110
115
W
e
i
g
h
t

(
%
)
Temperature (C)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(a)
0 200 400 600 800 1000
-2.0
-1.5
-1.0
-0.5
0.0
H
e
a
t

F
l
o
w

(
W
/
g
)
Temperature (C)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(b)
Fig. 2. Sulfated zirconia prepared using sulfuric acid with different digestion times
(a) TGA (b) DSC.
loss stages were observed in the TGA prole with rst one located
below150

Candis associatedwitha strong endothermic peakcen-


tered at about 80

C. This weight loss is attributed to the removal


of the adsorbed water on the surface. Second stage of weight loss is
between 150

C and 500

C corresponding to the removal of ter-


minal hydroxyl groups bonded to the surface of zirconia. Strong
exothermic peaks were observed over the temperature range of
420830

C. In this region of the exothermic peaks, no weight loss


is observed in the TGA curve, and hence the exothermic peaks
are attributed to the transition from an amorphous to a tetrago-
nal metastable phase of zirconia, i.e., a topotatic crystallization of
tetragonal zirconia on nuclei present in the amorphous phase. The
centers of these exothermic peaks for samples prepared at 0, 1,
3, 6, 12, 24, and 48h of digestion times were located at 455, 464,
512, 563, 635, 708, and 779

C, respectively, clearly indicating the


increase in crystallization temperature with increase in digestion
time. This is attributed to the changes in the as-prepared materials
particle sizes which have been observed to decrease with increase
in digestion time [16]. It can also be seen that the more the diges-
tion time, the lesser the weight loss (results not shown) due to the
removal of hydroxyl groups through the process of polymerization
between hydrous zirconia particles [21]. The thermo-gravimetric-
calorimetric analysis of the SZ and CSZ are shown in Figs. 2 and 3,
respectively. Similar to the hydrous zirconia samples, the weight
V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206 199
0 200 400 600 800 1000
40
60
80
100
120
W
e
i
g
h
t

(
%
)
Temperature (C)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(a)
0 200 400 600 800 1000
-2.5
-2.0
-1.5
-1.0
-0.5
0.0
0.5
H
e
a
t

F
l
o
w

(
W
/
g
)
Temperature (C)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(b)
Fig. 3. Sulfated zirconia prepared using chlorosulfonic acid with different digestion
times (a) TGA (b) DSC.
loss until 500

C is attributed to the removal of moisture and


adsorbed species. The second part of the TGA prole shows that
there is a steep decrease in the weight in the temperature range
of 580795

C. It was also observed that the more the digestion


time the higher the temperature needed to start this weight loss.
This weight loss is associated with an exothermalong with a small
endotherm, whose intensity and the position was also found to
be dependent on the time of digestion. The exotherm indicating
the crystallization temperature for SZ samples is clearly seen to be
shifted toward higher temperature compared to zirconia. The shift
in crystallization temperature was observed to be maximum for
sample with no digestion (453624

C) and then decreased with


increase in digestion time (463649

C for 1h, 464686

C for 3h
and 635701

C for 12h digestion time), and was almost none for


sample digested for 48h. The intensity of the peak was also seen to
be higher for the lowdigestion time samples compared to the ones
digested for longer times.
The endothermassociated with the weight loss is attributed to
the decomposition of sulfate species on the zirconia surface, which
leads to the formation of SO
3
moieties [22]. The temperature of
sulfates decomposition is also observed to be dependent on the
time of digestion, attributed to the change in the sulfate state on
the zirconia surface (different co-ordination with ZrO
2
sites) [23].
The DSC prole of CSZ in Fig. 3b shows an additional endotherm
in the temperature range of 220330

C which is attributed to the


removal of organic additives especially ethylene dichloride used
as solvent for chlorosulfonic acid. Interestingly, no exotherm was
observed for the crystallization whereas an endotherm showing
the decomposition of sulfate species was clearly seen. It is believed
that the exothermic crystallization of the zirconia may occurs
just before or with the endothermic decomposition of the sulfate
species and the shape of TGADSC prole depends on the relative
heat balance between these exothermicendothermic transforma-
tions [24]. Thus, the absence of the exothermic peak in the case of
CSZ could be attributed to the more dominant endothermic heat of
sulfate ions decomposition than the exothermic heat of crystalliza-
tion could be due to the presence of more sulfate ions compared to
SZ samples.
3.2. X-ray diffraction
The surface acidity and subsequently the catalytic activity
of the sulfated zirconia have been found to be dependent on
the calcination temperature of sulfated zirconia as well as the
pre-sulfonation drying temperature of zirconium hydroxide. The
zirconium hydroxide needs to be in the amorphous phase before
the sulfonation. Furthermore, the calcinations temperature of sul-
fated zirconia should yield crystalline tetragonal phase without the
degradation of much of the sulfate groups. Comelli et al. [25] found
that sulfated zirconia calcined between 530 and 605

C showed
highest catalytic activity for n-hexane isomerization reaction. A
high sulfur concentration and co-existence of S
4+
and S
6+
over
amorphous material was observedfor thesulfatedzirconiacalcined
at temperature below 500

C. Whereas for calcination tempera-


ture above 500

C, decreased sulfur concentration and existence of


only S
6+
on tetragonal zirconia structure was detected, which was
attributed to the higher catalytic activity for isomerization reac-
tion. In addition to the S
6+
inuence, the higher catalytic activity
of sulfated zirconia calcined at higher temperature was suggested
to be due to the formation of active sites by effectively binding
sulfate groups to the zirconia surface and also partial removal of
sulfate species from highly uncoordinated sites on ZrO
2
, creat-
ing strong Lewis acid sites upon calcination at higher temperature
[26,27]. Many researchers have suggested that 600650

C tem-
peratures are favorable for the formation of highly active sulfated
zirconia catalyst [11,2830]. Fig. 4 shows the XRD patterns of the
sulfated zirconia prepared from zirconium hydroxide synthesized
withdigestiontimes rangingfrom0hto48h, usingsulfuric acidand
chlorosulfonic acid as two different sulfonating agents; calcined at
600

C and 650

C temperature. When sulfuric acid was used, cal-


cination at 600

C temperature formed crystallites with tetragonal


structure for 0h and 1h digestion time. For higher digestion times
the materials is only partially crystallized to tetragonal structure
showing a broad peak at 30.2

2 value. Upon calcination at 650

C,
material was found to be crystallized with 100% tetragonal phase
structure, however lower crystallinity was observed for materials
digested for 3h and 6h.
When chlorosulfonic acid was used, 100% tetragonal structure
was obtained upon calcination at 600

C and with further increase


in calcination temperature to 650

C, material retained tetragonal


structure. It was also observed that for 0h digestion sample no
monoclinic phase was formed in both the cases even after calci-
nation at 650

C temperature. However, in an earlier studies we


have seenthat the zirconia preparedwithno digestionandcalcined
at 600

C consisted of a mixture of tetragonal (52%) and monoclinic


phases (48%) [16]. This indicates that the presence of sulfate groups
does stabilize the tetragonal structure of zirconia due to the strong
interactions between the sulfate and ZrO
2
suppressing or delaying
the phase transition [31,32]. The sizes of the crystallites calculated
using Scherrer equation were in the range of 620nm depending
200 V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206
20 30 40 50 60 70
0
500
1000
1500
2000
2500
3000
3500
4000
4500
5000
5500
I
n
t
e
n
s
i
t
y
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(b) CSZ
20 30 40 50 60 70
500
1000
1500
2000
2500
I
n
t
e
n
s
i
t
y
2 2
2 2
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(a) SZ
20 30 40 50 60 70
0
3000
6000
9000
12000
15000
18000
21000
I
n
t
e
n
s
i
t
y
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(c) SZ
20 30 40 50 60 70
2000
4000
6000
8000
10000
I
n
t
e
n
s
i
t
y
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(d) CSZ
Fig. 4. X-ray diffraction pattern of sulfated zirconia prepared with different digestion times and calcined at 600

C (a) SZ (b) CSZ and at 650

C (c) SZ (d) CSZ.


uponthe digestiontime, sulfonating agent andthe calcinationtem-
perature.
3.3. FTIR studies
Figs. 5 and 6 showthe FTIR spectra for the SZ and CSZ prepared
with different digestion times, respectively. FTIR spectra for both
SZ and CSZ showed the hydration of all the samples giving rise to a
strong, broad, and unresolved band in the 36003000cm
1
region
(see Supplemental information), assigned to physisorbed and pos-
sibly coordinated water, accompanied by a broad band close to
1620cm
1
that is ascribed to the bending mode (
HOH
) of coordi-
nated molecular water associated with the sulfate group [3335].
The SZ dried at 110

C showed broad peak in the region


of 8001350cm
1
with four clear peaks at about 1202cm
1
,
1123cm
1
, 1043cm
1
and 983cm
1
; characteristic of inorganic
chelating bidentate sulfate ion co-ordinated to metal cation, and
which are assigned to asymmetric and symmetric stretching fre-
quencies of S OandS Obonds [27]. Uponcalcinationat 600

Cand
650

C, only a broad peak with small shoulders at about 1216cm


1
,
1133cm
1
and 1057cm
1
instead of clear peaks were observed
due to the decreased concentration of sulfate species due to the
decomposition at high temperature. The shifting of the sulfate
vibration band toward higher frequencies upon calcination could
be attributed to the formation of strong bonding between sulfates
and zirconia atoms. Similar observation was made for the CSZ with
110

C dried sample showing various peak (at about 1045cm


1
,
1105cm
1
, 1170cm
1
and 1280cm
1
) for S O and S O
vibrations as shown in Fig. 6. For 110

C dried CSZ an intense peak


was observed at 1620cm
1
attributed to HOHbond bending vibra-
tion for non digested sample and the intensity was observed to
decrease with increase in digestion time. In our earlier studies on
zirconia synthesis [16], a strong transmittance band at 935cm
1
which is attributed to the Si Hbending vibrations for non-calcined
samples and at 1050cm
1
attributed to the asymmetric stretching
vibrations of the Si O Si bond was observed. The interference of
this peak with the sulfate vibration peaks could also be the reason
for sulfate IR peaks disappearance. Corma and Garcia [36] observed
that the IR bands for sulfates in calcined sulfated zirconia cata-
lysts whenexposedto moisture were shownto disappear at certain
H
2
O:S molar ratio. The changes in IR spectra upon exposure to the
moisture were attributed to successive formation of H
2
SO
4
, HSO
4

and SO
4
2
fromwater sensitive SO
3
groups. This could also be one
of the reasons for the observed disappearance of sulfate peaks for
calcined SZ and CSZ samples in the present study.
3.4. Ammonia TPD studies
Temperature programmed desorption (TPD) of ammonia was
used to measure the total acid strength and the acid sites distribu-
tion on the surface of sulfated zirconia. Fig. 7 shows the acid sites
distribution for the SZ prepared from zirconium hydroxide syn-
thesized at 1, 3 and 6h and calcined at 600

C and 650

C. It was
observed that for all three digestion times and for both calcination
temperatures, SZ exhibit two peaks, at about 300

C and 600

C.
The CSZ calcined at 600

C, exhibited 4 different peaks located


V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206 201
Table 1
Textural properties of the SZ and CSZ calcined at 600

C temperature.
Material Digestion time
(h)
Surface area
(m
2
/g)
Pore volume
(cm
3
/g)
Avg. pore size
(nm)
Zirconia 3 141.6 0.1415 3.996
SZ 1 114.1 0.0965 3.381
3 149.0 0.1306 3.507
6 165.5 0.1492 3.606
12 158.8 0.1546 3.892
24 177.6 0.1845 4.157
48 176.5 0.1807 4.096
CSZ 1 18.02 0.0664 14.74
3 20.93 0.0778 14.71
6 23.73 0.0881 14.85
12 18.90 0.0516 10.91
24 20.98 0.0523 9.982
48 20.07 0.0478 9.528
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.75
0.80
0.85
0.90
0.95
1.00
1.05
1.10
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(a)
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.90
0.95
1.00
1.05
1.10
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(b)
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.80
0.85
0.90
0.95
1.00
1.05
1.10
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(c)
Fig. 5. FTIR spectra of sulfated zirconia prepared using sulfuric acid with different digestion times and calcined at (a) 110

C (b) 600

C (c) 650

C.
in the temperature range of 156200

C, 280300

C, 380400

C,
and 560600

C as shown in Fig. 8. However, upon calcination


at 650

C the two intermediate peaks merged together to give


one broad peak. It was also observed that for both SZ and CSZ,
the intensity of peaks decreased when calcination temperature
increased from600

C to 650

C due to the decomposition of sulfate


groups as showninthe TGADSCanalysis. The ammonia desorption
peak at temperature below 200

C belongs to the physisorp-


tion/chemisorptions of ammonia molecules on weak acidic sites.
The peaks at about 300

C and 480

C show the existence of inter-


mediate strength acidic sites and nally the peak at 560600

C
demonstrates the presence of superacidic sites on the surface of
zirconia [14]. By comparing Fig. 7 and Fig. 8, it can also be observed
that the CSZ possesses more superacidic sites compared to the SZ.
The quantitative measurement of total acid sites for the SZ and CSZ
prepared at various digestion times and calcination temperature
is depicted in Fig. 9. It is observed that, for both SZ and CSZ, the
total number of acids sites initially increased when digestion time
202 V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.7
0.8
0.9
1.0
1.1
1.2
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(a)
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.75
0.80
0.85
0.90
0.95
1.00
1.05
1.10
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(b)
2400 2200 2000 1800 1600 1400 1200 1000 800 600
0.80
0.85
0.90
0.95
1.00
1.05
1.10
T
r
a
n
s
m
i
t
t
a
n
c
e

(
%
)
Wavenumber (cm
-1
)
48 h
24 h
12 h
6 h
3 h
1 h
0 h
(c)
Fig. 6. FTIR spectra of sulfated zirconia prepared using chlorosulfonic acid with different digestion times and calcined at (a) 110

C (b) 600

C (c) 650

C.
increased from0 to 3h but then observed to decrease with further
increase in digestion time to 24h. Upon more increase in diges-
tion time to 48h, the acid sites were seen to increase again by a
small number. Similar observation was made about the XRD peak
intensities for the sulfated zirconia samples prepared at different
digestion times (Fig. 4c and d) wherein lower XRD peak intensities
were observed for 3h and 6h digested samples. These observa-
tions can be explained on the basis of the effect of digestion time on
surfaceareaandhydroxyl groups onthesurfaceof hydrous zirconia.
Chen et al. [37] demonstrated the importance of surface OH
groups for the formation of sulfated zirconia. They suggested that
the formation of sulfated zirconia is a two step process. In the
rst step sulfate group displaces the surface OH group and upon
calcination acid sites are formed through oxolation process form-
ing stronger chemical bonds [38]. In an earlier work [16], we
demonstrated that the process of digestion leads to an extensive
polymerization through the condensation of the hydroxyl groups
in the hydrous zirconia to formordered three-dimensional porous
structure with greater thermal stability. Thus an increase in diges-
tion time inuences both the surface hydroxyl groups as well as the
surface area of the material. As the digestiontime increases the sur-
face areas increases at the expense of surface OH groups which are
important for the anchoring of sulfate groups during the sulfona-
tion process. On the other hand, more surface area provides more
distributed anchoring sites for unhindered attachment of sulfate
groups. Thus, with lower digestion time we have more OH groups
but low surface area and at higher digestion times we have less
OH groups but higher surface area, which explains the observation
made about the variations in crystallinity in the XRD analysis and
the existence of optimum number of acid sites for 3h digestion
time.
3.5. Nitrogen adsorptiondesorption
The nitrogen adsorptiondesorption isotherms for the zirconia,
SZ and CSZ prepared with 3h digestion time and calcined at 600

C
temperature are shown in Fig. 10. The isotherms resemble the type
IV isotherms with hysteresis loop of type H2 based on IUPAC clas-
sication [39]. The hysteresis loop is associated with the capillary
condensation taking place in the mesopores signifying existence of
mesoporous structure in the calcined sulfated zirconia. The type H2
hysteresis loop is attributed to the ink-bottle shaped pores (pores
with narrownecks and wide bodies). Table 1 presents the values of
surface area, pore volume and average pore sizes of the SZ and CSZ
prepared with various digestion times calcined at 600

C tempera-
ture. A slightly higher surface area was observed upon sulfonation
using sulfuric acid compared to the non-sulfated zirconia. This sup-
ports the observation made in the TGADSC studied wherein delay
in the crystallization temperature was observed upon sulfonation.
The surface area was also observed to increase with increase in
digestion time. Several reports have suggested that the introduc-
tion of sulfate anions disturbs the transition of amorphous phase to
V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206 203
-200 0 200 400 600 800 1000 1200 1400 1600 1800 2000
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
(a) SZ-1_600
0 200 400 600 800 1000 1200 1400 1600 1800
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(b) SZ-1_650
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 400 800 1200 1600 2000
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(c) SZ-3_600
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 200 400 600 800 1000 1200 1400 1600 1800
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(d) SZ-3_650
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 400 800 1200 1600 2000
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(e) SZ-6_600
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 200 400 600 800 1000 1200 1400 1600 1800
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(f) SZ-6_650
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
Fig. 7. NH
3
TPD of sulfated zirconia prepared using sulfuric acid and zirconia prepared with digestion times and calcination temperature of (a) 1h, 600

C (b) 1h, 650

C (c)
3 h, 600

C (d) 3h, 650

C (e) 6h, 600

C (f) 6h, 650

C, respectively.
crystalline phase and the extent of delay in transition depends on
the concentration of active sulfate species on the zirconia surface
[13]. Thus it can be concluded that the time of digestion certainly
has the inuence on the number of active sulfate sites on the zir-
conia surface. A drastic reduction in the surface area and pore
volume and increase in the average pore size was observed upon
sulfonation using chlorosulfonic acid. The shift of the hysteresis
loop toward the higher relative pressures signies the formation of
bigger pores at expense of breaking of smaller pores. This could be
due to the destruction of mesoporous structure of hydrous zirco-
nia caused by the highly corrosive action of the chlorosulfonic acid
[32,40].
3.6. Catalysts performance for esterication reaction
The catalytic activity of the prepared SZ and CSZ catalysts syn-
thesized with digestion time of 1, 3, 6, 12, 24 and 48h and calcined
at 600

C temperature were tested for the esterication of oleic


acid in the soybean oil. The conversion was calculated using the
following equation;
X
A
=
AV
0
AV
AV
0
(2)
where AV
0
is initial acid value and AV is acid value at any time
t. As the purpose of the study was to investigate the effect of
204 V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206
0 200 400 600 800 1000 1200 1400 1600
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(b)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 200 400 600 800 1000 1200 1400 1600 1800
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
(a)
0 400 800 1200 1600 2000 2400
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
(c)
0 200 400 600 800 1000 1200 1400 1600 1800
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(d)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
-200 0 200 400 600 800 1000 1200 1400 1600 1800 2000
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
(e)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
0 200 400 600 800 1000 1200 1400 1600
100
200
300
400
500
600
Temperature
Signal
Time (sec)
T
e
m
p
e
r
a
t
u
r
e

(
o
C
)
0
10
20
30
40
50
60
70
T
C
D

S
i
g
n
a
l

(
m
V
)
(f)
Fig. 8. NH
3
TPDof sulfated zirconia prepared using chlorosulfonic acid and zirconia prepared with digestion time and calcination temperature of (a) 1h, 600

C (b) 1h, 650

C
(c) 3h, 600

C (d) 3h, 650

C (e) 6h, 600

C (f) 6h, 650

C, respectively.
catalyst preparation parameters on the rate of reaction and %
conversion, all reaction parameters were kept constant at 60

C
reaction temperature, methanol/oil molar ratio of 9:1 and 2% cat-
alyst concentration. The leaching test of the catalyst was carried
out in methanol and also in the reaction mixture (without oleic
acid) by following the procedure reported by Suwannakarn et al.
[41]. It was observed that both SZ and CSZ did not show any
leaching into the methanol or reaction mixture. Figs. 11 and 12
summarize the results obtained for catalytic activity of SZ and CSZ
in the esterication reaction. The results suggest that the type of
sulfonating agent has strong inuence on the catalytic activity of
the sulfated zirconia. In the case of SZ, maximum conversion of
22.49% was obtained at the end of 80min whereas for the same
reaction time it was 84.33% conversion for CSZ. The signicantly
higher activity of the CSZ even after having relatively lower sur-
face area compared to the SZ is attributed to existence of more
superacidic sites on the surface of CSZ as discussed earlier, hence,
resulting in higher conversion. The digestion time of zirconia was
also observed to inuence the rate of reaction and the nal con-
version of oleic acid for both SZ and CSZ. The sulfated zirconia
prepared from hydrous zirconia synthesized at 3h digestion time
showed higher esterication activity compared to the hydrous
zirconia synthesized at other digestion times. These results con-
rm the observance of more number of total acidic sites for 3h
V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206 205
Fig. 9. Variation of total number of acid sites (mol/g) with respect to digestion
time and sulfonating agent on the surface of sulfated zirconia calcined at 600

C and
650

C temperature.
0.0 0.2 0.4 0.6 0.8 1.0
0
10
20
30
40
50
60
70
80
90
100
110
120
130
140
150
SZ
CSZ
D
A
V
o
l
u
m
e

[
c
c
/
g
]
P/Po
Zirconia
Fig. 10. Nitrogenadsorptiondesorptionisothermfor zirconia, SZ and CSZ prepared
at 3h digestion time. A: adsorption, D: desorption.
Fig. 11. Conversion vs time plot for SZ prepared at different digestion times for
oleic acid estericationreaction. Methanol/acid oil: 9/1, temperature: 60

C, catalyst
loading: 2% (wt % of acid oil).
Fig. 12. Conversion vs time plot for CSZ prepared at different digestion times for
oleic acid estericationreaction. Methanol/acid oil: 9/1, temperature: 60

C, catalyst
loading: 2% (wt % of acid oil).
digested sample compared to 1 and 6h which is demonstrated in
Fig. 9.
4. Conclusions
The hydrous zirconia prepared with different hydrothermal
treatment times was used to synthesize the sulfated zirconia solid
acidcatalyst usingsulfuric acidandchlorosulfonic acids as sulfonat-
ing agents. The digestion time and the type of sulfonating agents
were found to have signicant inuence on structural, textural and
catalytic properties of the prepared sulfated zirconia catalyst. The
sample prepared with 3h digestion time was observed to possess
the highest number of total acid sites. For the esterication of
oleic acid in soybean oil, the CSZ catalyst prepared with 3h diges-
tion time and 600

C calcination temperature showed the highest


catalytic activity of about 85% conversion in just 80min at 60

C
reaction temperature. The higher catalytic activity of CSZ though
with signicantly lower surface area compared to SZ is attributed
to the presence of higher surface superacidic sites compared with
SZ.
Acknowledgements
The authors acknowledge the funding received from National
Science Foundation (NSF) for the nancial assistance via Award
CBET 0651811. The authors are also grateful to the College of
Engineering at North Carolina Agricultural and Technical State Uni-
versity(NCAT) for partial support for this project. The authors thank
the Center for Advanced Materials and Smart Structures (CAMSS)
at NCAT for the use of their XRD for material characterization.
Appendix A. Supplementary data
Supplementary data associated with this article can be
found, in the online version, at http://dx.doi.org/10.1016/
j.apcata.2013.05.005.
References
[1] A. Demirbas, Prog. Energy Combust. Sci. 31 (2005) 466487.
[2] M.S. Graboski, R.L. McCormick, Prog. Energy Combust. Sci. 24 (1998) 125164.
[3] K. Bozbas, Renew. Sust. Energy Rev. 12 (2008) 542552.
[4] F. Ma, M.A. Hanna, Biores. Technol. 70 (1999) 115.
[5] I. Atadashi, M. Aroua, A. Abdul Aziz, N. Sulaiman, Renew. Sust. Energy Rev. 16
(2012) 32753285.
[6] B.R. Moser, In Vitro Cell Dev. Biol.-Plant 45 (2011) 229266.
[7] E. Lotero, Y. Liu, D.E. Lopez, K. Suwannakarn, D.A. Bruce, J.G. Goodwin Jr., Ind.
Eng. Chem. Res. 44 (2005) 53535363.
206 V.G. Deshmane, Y.G. Adewuyi / Applied Catalysis A: General 462463 (2013) 196206
[8] M.E. Borges, L. Daz, Renewable, Sustain. Energy Rev. 16 (2012) 28392849.
[9] K. Nuithitikul, J. Limtrakul, Int. J. Chem. React. Eng. 10 (2012) 129.
[10] A.A. Kiss, A.C. Dimian, G. Rothenberg, Adv. Synth. Catal. 348 (2006) 7581.
[11] G.D. Yadav, J.J. Nair, Microporous Mesoporous Mater. 33 (1999) 148.
[12] B.M. Reddy, M.K. Patil, ChemInform40 (2009) 21862208.
[13] T. Yamaguchi, K. Tanabe, K. Yao Chin, Mater. Chem. Phys. 16 (1987) 6777.
[14] A. Corma, V. Fornes, M. Juan-Rajadell, J. Nieto, Appl. Catal. A: Gen. 116 (1994)
151163.
[15] D. Farcasiu, J.Q. Li, Appl. Catal. A: Gen. 128 (1995) 97105.
[16] V.G. Deshmane, Y.G. Adewuyi, Microporous Mesoporous Mater. 148 (2011)
88100.
[17] G.D. Yadav, A.D. Murkute, J. Catal. 224 (2004) 218223.
[18] L. DSouza, A. Suchopar, K. Zhu, D. Balyozova, M. Devadas, R.M. Richards, Micro-
porous Mesoporous Mater. 88 (2006) 2230.
[19] E.P. Barrett, L.G. Joyner, P.P. Halenda, J. Am. Chem. Soc. 73 (1951) 373380.
[20] A.L. Patterson, Phys. Rev. 56 (1939) 978982.
[21] A. Cleareld, Inorg. Chem. 3 (1964) 146148.
[22] Q.H. Xia, K. Hidajat, S. Kawi, J. Catal. 205 (2002) 318331.
[23] M. Hino, M. Kurashige, H. Matsuhashi, K. Arata, Thermochim. Acta 441 (2006)
3541.
[24] R. Srinivasan, R.A. Keogh, D.R. Milburn, B.H. Davis, J. Catal. 153 (1995) 123130.
[25] R.A. Comelli, S.A. Canavese, S.R. Vaudagna, N.S. Fgoli, Appl. Catal. A: Gen. 135
(1996) 287299.
[26] X. Li, K. Nagaoka, R. Olindo, J.A. Lercher, J. Catal. 238 (2006) 3945.
[27] T. Yamaguchi, T. Jin, K. Tanabe, J. Phys. Chem. 90 (1986) 31483152.
[28] C. Guo, S. Yao, J. Cao, Z. Qian, Appl. Catal. A: Gen. 107 (1994) 229238.
[29] C. Morterra, G. Cerrato, M. Signoretto, Catal. Lett. 41 (1996) 101109.
[30] M. Tran, N. Gnep, G. Szabo, M. Guisnet, Appl. Catal. A: Gen. 171 (1998) 207217.
[31] J.R. Sohn, H.W. Kim, J. Mol. Catal. 52 (1989) 361374.
[32] J. Laizet, A. Soiland, J. Leglise, J. Duchet, Top. Catal. 10 (2000) 8997.
[33] D.A. Ward, E.I. Ko, J. Catal. 150 (1994) 1833.
[34] D. Sarkar, D. Mohapatra, S. Ray, S. Bhattacharyya, S. Adak, N. Mitra, Cer. Int. 33
(2007) 12751282.
[35] F. Babou, G. Coudurier, J.C. Vedrine, J. Catal. 152 (1995) 341349.
[36] A. Corma, H. Garcia, Catal. Today 38 (1997) 257308.
[37] F.R. Chen, G. Coudurier, J.F. Joly, J.C. Vedrine, J. Catal. 143 (1993) 616626.
[38] C.R. Vera, J.M. Parera, J. Catal. 166 (1997) 254262.
[39] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603619.
[40] J. Mass, A.G. Mayer, Chlorosulfonic Acid in Ullmanns Encyclopedia of Industrial
Chemistry, Wiley, NewYork, USA, 2005.
[41] K. Suwannakarn, E. Lotero, J.G. Goodwin Jr., C. Lu, J. Catal. 255 (2008) 279286.

Você também pode gostar