Sie sind auf Seite 1von 9

View Online / Journal Homepage

Catalysis
Science & Technology

Dynamic Article Links

Cite this: DOI: 10.1039/c2cy20462b

PAPER

www.rsc.org/catalysis

Sulfated zirconia: an ecient solid acid catalyst for esterication of


myristic acid with short chain alcoholsw

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

K. Saravanan, Beena Tyagi* and H. C. Bajaj


Received 11th June 2012, Accepted 20th July 2012
DOI: 10.1039/c2cy20462b
Sulfated zirconia (SZ) catalysts prepared by a two-step solgel method and calcined at 600700 1C
were evaluated for esterication of myristic acid with methanol using varied acid to alcohol ratio,
reaction temperature and catalyst concentration. An exceptionally small concentration of SZ
catalysts (0.1250.5 wt% to acid) exhibited 98100% conversion of myristic acid with methanol at
60 1C after 5 h. The conversion was decreased with an increase in the alkyl chain of alcohol from
methanol to butanol, however, similar conversion was achieved by increasing the reaction
temperature to 90 1C. The ester formation was selective, irrespective of alcohol and other reaction
variables. The calcination temperature has strong inuence on the structural, textural and acidic
features and thus on the activity and re-usability of SZ catalysts. The reaction is sensitive to
moisture present in methanol or reaction mixture. The studied reaction is Bronsted acid catalyzed
and the SZ catalyst having higher number of Bronsted acid sites was re-used successfully without
signicant loss in activity; whereas the SZ catalyst having lower number of Bronsted acid sites
showed a decrease in activity (B28%) after ve reaction cycles. The results clearly indicated the
necessity of higher number of Bronsted acid sites for better performance and recycling of the SZ
catalysts for esterication of myristic acid with methanol under the conditions studied.

1. Introduction
Myristic acid is a saturated fatty acid (n-tetradecanoic acid,
C14H28O2) naturally occurring in nutmeg Myristica fragrans.
It is also found in palm kernel oil, coconut oil, butter fat and
animal fats, which are used to produce fatty acid alkyl esters
(FAAEs), used as biodiesel.1 It increases low density lipoprotein
cholesterol making it one of the most hypercholesterolemic of
the saturated fatty acids.2 Alkyl myristate of short chain
alcohols namely methanol, ethanol, propanol and butanol
are used in perfumes, avors, cosmetics, personal care
products, food additives, detergents, soaps, lubricants for
textiles, plasticizers etc.3 Besides, methyl myristate has been
approved as a new active constituent in veterinary chemical
products by Australian pesticides and veterinary medicines
authority4 and ethyl myristate is used as a marker of excessive
ethanol consumption that can be isolated from the hair of an
alcoholic individual.5

Discipline of Inorganic Materials and Catalysis, Council of Scientic


and Industrial Research (CSIR), Central Salt and Marine Chemicals
Research Institute (CSMCRI), G. B. Marg, Bhavnagar, Gujarat 364 002,
India. E-mail: btyagi@csmcri.org; Fax: +91-278-2566970
w Electronic supplementary information (ESI) available: PXRD pattern,
TG/DTA, nitrogen sorption isotherms, IR spectra, DRIFT spectra from
RT to 450 1C, pH vs. time after stirring the SZ-600 catalyst in water and
methanol and comparison with other acid catalysts. See DOI: 10.1039/
c2cy20462b

This journal is

The Royal Society of Chemistry 2012

Conventionally, FAAEs have been prepared by homogeneous acid catalyzed esterication of the corresponding acid
with an alcohol or homogeneous base catalyzed transesterication of oils and fats. Industrial esterication processes are
carried out in the presence of homogeneous Bronsted acid
catalysts such as sulfuric, p-toluene sulfonic or phosphoric
acid.6 To overcome the disadvantages associated with homogeneous catalysts, a number of solid acid catalysts such as
inorganic metal oxides, heteropoly acids and sulfonic acid
based resins etc. are being studied to produce FAAEs via
esterication of fatty acids (FAs) or transesterication of oil
with alcohols.6,7 Among metal oxide based solid acid catalysts,
sulfated zirconia (SZ) has exhibited signicant activity in both
the reactions.613 Kiss et al.9 have screened various solid acids
including zeolites, ion exchange resins and metal oxides for the
esterication of lauric acid with dierent alcohols and found
SZ as the most active solid acid catalyst; zeolites were not
suitable due to their microporous nature having diusion
limitations of large fatty acid molecules, whereas, ion
exchange resins have low thermal stability. SZ is a potential
solid acid catalyst for alkane isomerization at mild temperature14
and many other commercially important organic transformations due to its strong acid properties.1517
Few studies have been reported for the esterication of
myristic acid using SZ18,19 and niobia20 metal oxide catalysts.
Zeolites21 and acidic ion exchange resins namely Amberlyst-15
and silica based resin having sulfonic acid groups22 were not
Catal. Sci. Technol.

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

found eective as compared with metal salts or homogenous


p-TSA. Among various metal salts, zirconium sulfate showed
signicant activity.21 MCM-48 supported tungstophosphoric
acid23 and ZrOCl28H2O24 have been studied with long chain
alcohols such as cetyl alcohol in sc-CO2 and other organic
solvents. However, most of these studies have been done for
various long chain saturated and unsaturated FAs or along
with triglycerides at high temperature (120180 1C)18,19,23,24
in the presence of higher concentration of solid catalysts
(515 wt%).1921 To the best of our knowledge, no detailed
study has been reported for the esterication of myristic
acid with methanol and other short chain alcohols over SZ
catalysts.
We have studied SZ catalysts for various reactions mainly
acetylation,25a,b isomerization,25c,d Pechmann,25e,f condensation26a
and recently esterication of caprylic acid.26b Herein, we
report a systematic detailed study of the esterication of
myristic acid with methanol at lower temperature in the
presence of small concentration of SZ catalysts. The eect of
various reaction parameters such as acid to alcohol ratio,
catalyst concentration and reaction temperature along with
other short chain alcohols namely ethanol, n-propanol and
n-butanol has been investigated. The eect of calcination
temperature on the acidity and activity, re-usability and
deactivation of SZ catalysts has also been addressed. In our
earlier report25c we have observed that selectivity of the
isomerized products can be varied depending upon the acidity
of SZ catalysts in terms of Bronsted (B) to Lewis (L) acid site
concentration ratio. Herein, we have found that the acidic
properties in terms of Bronsted acid site concentration, B/L
ratio and total surface acidity play a major role in the activity
and deactivation of SZ catalysts for esterication of myristic
acid with methanol under the conditions studied. The present
study gives novel insight into the inuence of the acidic
properties on the catalytic performance of fresh and re-used
SZ catalysts and the use of a very small amount of the catalyst
resulting in maximum conversion of acid at lower temperature.

2. Results and discussion


2.1

Catalyst characterization

The partial characterization of SZ catalysts has been reported


in our previous paper.26a In the context of the present work
few salient features are mentioned briey to understand the
structureactivity relationship.
Table 1

Fig. 1 TEM image of the SZ-600 catalyst at low resolution (inset)


high resolution.

A powder X-ray diraction (PXRD) pattern of SZ catalysts


showed the presence of tetragonal crystalline phase after
calcination at 600700 1C. SZ-600 was less crystalline and
the crystallinity of the samples increased with increasing
calcination temperature (Fig. S1, ESIw). The catalysts were
found to have nano-crystallite size in the range of 1116 nm
(Table 1). The bulk sulfur content before calcination was
3.7 wt%, which successively decreased after calcination at
600700 1C (Table 1). The TG/DTA prole (Fig. S2, ESIw)
of the catalyst also conrmed the crystallization of the SZ
sample at 4600 1C.
A TEM micrograph of SZ-600 (Fig. 1) showed the aggregates of SZ crystallites varying from 6.9 to 10.2 nm with an
average crystallite size of 7.3 nm, which was in agreement with
crystallite size calculated by PXRD. The high-resolution TEM
image (inset, Fig. 1) exhibited the lattice fringes, the width
between two fringes (B2.86 A) also agreed with d-spacing
(B2.93 A) of characteristic peaks of tetragonal phase of
zirconia (2y = 30.22) shown by PXRD.
BET surface area and pore volume were found to decrease
with an increase in calcination temperature; however SZ-700
showed higher pore diameter due to collapsing of pore walls at
higher temperature (Table 1). The N2 adsorption isotherms
(Fig. S3, ESIw) showed the increase in adsorption at higher
relative pressure indicating the presence of larger size of mesopores in the samples.27 The inection point at P/Po = B0.40.5

Characterization of sulfated zirconia catalysts


DRIFTb

NH3-TPD

Crystallite Sulfura SBET


Average pore
BJH pore
Cyclohexanol Temp. Acid sites Total acid sites B acid site B/L
Catalyst sizea (nm) (wt%) (m2 g 1) volume (cm3 g 1) diameter (nm) conversion (%) (1C)
(mmol g 1) (mmol g 1)a
(%T cm 1) ratio
SZ-600
SZ-650

11
15

2.6
1.04

75
74

0.12
0.09

6.5
5.7

91
87

SZ-700

16

0.87

54

0.08

7.1

85

Data from ref. 26a.

Catal. Sci. Technol.

719
116
564
112
400
627
793

2.50
0.80
0.19
0.38
0.07
0.14
0.08

2.5
0.99

3690
2939

2.07
1.66

0.67

2738

1.21

B and L = Bronsted and Lewis acid sites, at 150 1C.

This journal is

The Royal Society of Chemistry 2012

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

Fig. 2 Pore size distribution proles of SZ catalysts calcined at


dierent temperatures.

was not sharp, which reected that the pores have varied pore
size distribution (Fig. 2).
IR-spectra of the catalysts showed the presence of sulfate
groups in the range of 1245900 cm 1 (Fig. S4, ESIw). SZ-600
showed the presence of a broad intense peak showing
shoulders at 1230, 1138, 1040 and 990 cm 1, however, highly
crystalline SZ-650 and SZ-700 clearly exhibited the peaks at
12301245, 11361142, 10471056 and 988 cm 1, which are
assigned to asymmetric and symmetric stretching frequencies
of S]O and SO bonds and are characteristic of inorganic
chelating bidentate sulfate. The partially ionic nature of the
S]O bond is responsible for the Bronsted acid sites in sulfated
zirconia.28 The intensity of the sulfate peaks was reduced in
SZ-700 due to loss of sulfate species at higher calcination
temperature.
The dehydration of cyclohexanol over SZ-600 showed 91%
conversion of cyclohexanol with selective formation of cyclohexene indicating the presence of higher Bronsted acidity in
SZ-600, which slightly decreased in SZ-650 and SZ-700
(Table 1). Similarly, the total surface acidity analyzed
by NH3-TPD showed the highest number of acid sites
(2.5 mmol g 1) in SZ-600, which decreased ominously in
SZ-650 and SZ-700 (Table 1). The acidity in sulfated zirconia
is attributed to the presence of sulfate groups; as SZ-600 has
the highest sulfur content, it showed the highest number of
acid sites, the decrease in sulfur content with increasing
calcination temperature resulted in a decrease in acidity.
Furthermore, NH3 desorption occurred at relatively higher
temperature in the SZ-600, indicating the presence of only
strong acid sites (719 1C), whereas SZ-650 and SZ-700 showed
the presence of weak (112116 1C), moderate (400500 1C) to
strong (4600 1C) acid sites (Table 1).
The DRIFT spectra of pyridine adsorbed SZ catalysts
exhibited the characteristic peaks of pyridinium ions (Bronsted
acid sites) at 1540 cm1 and of covalently bonded pyridine
(Lewis acid sites) at 1441 cm1 (Fig. 3) along with peaks at
1616 cm1 and 1488 cm1 representing the Bronsted and total
acid sites respectively.25e,29 Both acid sites were observed to be
strong enough as they were present even after heating at 450 1C;
This journal is

The Royal Society of Chemistry 2012

Fig. 3 DRIFT spectra of SZ catalysts after pyridine desorption at


150 oC.

though the intensity of the peaks was decreased after successive heating (Fig. S5, ESIw). SZ-600 and SZ-650 showed
intense peaks for Bronsted acid sites, whereas, in SZ-700 the
peak for the Lewis acid site was intense. The Bronsted acid site
concentration and B/L ratio, calculated from the characteristic
peak area at 150 1C, were in the descending order from SZ-600
to SZ-700 (Table 1).
2.2 Catalytic activity
Initially, esterication of myristic acid was studied with
methanol over the SZ-600 catalyst to optimize acid to alcohol
ratio, catalyst concentration and temperature to achieve the
maximum conversion of myristic acid and selectivity for
methyl myristate.
2.2.1 Eect of acid to alcohol ratio and catalyst concentration. The reaction was scrutinized using an acid to methanol
molar ratio of 1 : 5 to 1 : 20 over catalyst concentration in the
range of 0.1250.5 wt% to acid. The varying concentration of
the catalyst (0.1250.5 wt%) resulted in 98100% conversion
of myristic acid having 100% selectivity for methyl myristate
using an acid to methanol molar ratio of 1 : 20 at 60 1C after
7 h (Table 2). By decreasing the acid to methanol molar ratio
to 1 : 10, the catalyst showed 47% conversion with 0.125 wt%
of catalyst, which enhanced upon increasing the catalyst
concentration. By further decreasing the acid to methanol
ratio to 1 : 5, 82% conversion was observed with 0.5 wt%
catalyst concentration. The results indicated the requirement
Table 2 Esterication of myristic acid with methanol over varied
concentrations of SZ catalystsa
Conversion (wt%)
Catalyst (wt%)

Methyl myristate (wt%)

0.5
0.25
0.125

498 (82*)
74
47

498
498
498

100
100
100

Reaction conditions: temperature = 60 1C; time = 7 h. A


acid : methanol = 1 : 10; B acid : methanol = 1 : 20; *acid : methanol =
1 : 5.

Catal. Sci. Technol.

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

of higher acid to alcohol ratio to obtain complete conversion


of FAs as also reported in the literature.6
Theoretically, esterication reaction requires one mole of
alcohol for one mole of acid; however, in practice, a molar
ratio higher than stoichiometric ratio is needed to complete the
reaction having acceptable reaction rates. The higher amount
of methanol shifted the equilibrium to the right side, thus
achieving the maximum conversion even with lower catalyst
amount (0.125 wt%). On the other hand, higher concentration
(0.5 wt%) of the catalyst increased the availability of acid sites,
which favour the accessibility of more number of reactant
molecules to these acid sites and thus enhanced the conversion
in the presence of lower acid to alcohol ratio. Though a very
high acid to alcohol ratio has been used to achieve maximum
acid conversion of FAs, e.g., 1 : 50,30 1 : 4031 to 1 : 10011 and
1 : 20032 in the presence of HPA, SZ and WZ catalysts,
respectively; lower amount of alcohol is more desirable in
the industrial synthesis of FAAEs, and thus we preferred an
acid to methanol ratio of 1 : 10 along with 0.5 wt% of catalyst
(protocol A) for rest of the reactions. However, for comparison, we have also studied few reactions with an acid to
methanol ratio of 1 : 20 in the presence of 0.125 wt% of
catalyst (protocol B). It is worth to be noted here that the
amount of SZ catalyst used was exceptionally small (0.1250.5
wt% to myristic acid, i.e., acid : SZ (wt/wt) = 800200 : 1) and
to the best of our knowledge, not reported earlier. Similar
good activity was also found for caprylic acid with 0.5 wt% of
catalyst.26b

Fig. 4 Conversion of myristic acid with time at (a) 50 1C and


(b) 60 1C, (c) eect of fresh methanol (at 60 1C). Reaction conditions:
acid : methanol = 1 : 10; SZ = 0.5 wt%.

To optimize the reaction temperature, we have studied the


reaction kinetics both at 50 and 60 1C. Though the nal
conversion of myristic acid was similar (98100%) after 7 h
and remained steady afterwards both at 50 and 60 1C (Fig. 4),
the rate of increase in conversion was higher at 60 1C. Therefore, 60 1C temperature was chosen as optimum reaction
temperature for rest of the studies.

2.2.2 Eect of temperature. The reaction was carried out in


the temperature range of 4060 1C with both sets of reaction
variables as stated above. The reaction having (A) variables
showed 69% conversion of myristic acid at room temperature
(32 1C) after 7 h, which increased at 50 1C and remained
constant upon further increasing the temperature (Table 3).
Whereas, the reaction having (B) variables showed a gradual
increase in the conversion from 81% to 98% by increasing the
temperature from 40 to 60 1C. The availability of higher
concentration of the catalyst in the (A) reaction signicantly
enhanced the interaction between the acid sites and the
reactants by increasing the temperature from 40 to 50 1C.
However, the lower concentration of the catalyst in the (B)
reaction was further diluted in the presence of higher amount
of methanol and thus resulted in a gradual enhancement in the
conversion with an increase in the temperature from 40 to 60 1C.

2.2.3 Eect of calcination temperature. To study the inuence of calcination temperature on the activity of SZ catalysts,
all three samples that calcined in the range of 600700 1C have
been evaluated under both the reaction variables as stated
above (Table 4). SZ-600 and SZ-650 catalysts resulted in
similar conversion under (A) reaction variables, however,
SZ-700 showed a signicant decrease in the conversion
(61%). The acidity of SZ catalysts is ascribed to the presence
of sulfate species which undergo thermal decomposition during the calcination and therefore the loss of sulfate species with
increasing calcination temperature resulted in the decrease in
the acidity and thus the activity of the catalyst. Though SZ650 has lower sulfur content, Bronsted acidity, total surface
acidity and weak to moderate acid sites as compared to SZ-600
(Table 1); its catalytic activity for the studied reaction was not
aected. It may be due to the presence of minimum required

Table 3 Esterication of myristic acid with methanol at dierent


reaction temperaturesa

Table 4 Esterication of myristic acid with methanol over SZ


catalysts calcined at dierent temperaturesa

Conversion (wt%)

Conversion (wt%)

Temperature (1C)

Catalyst

32
40
50
55
60

69
71
498
498
498

81
87
91
498

SZ-600
SZ-650
SZ-700
Zr-400
Zr-600

498
498
61
28
23

498
80
58

Reaction conditions: A acid : methanol = 1 : 10 and catalyst =


0.5 wt%; B acid : methanol = 1 : 20 and catalyst = 0.125 wt%;
Reaction Time = 7 h.

Catal. Sci. Technol.

Reaction conditions: temperature = 60 1C; time = 7 h. A


acid : methanol = 1 : 10 & catalyst = 0.5 wt%. B acid : methanol =
1 : 20 & catalyst = 0.125 wt%.

This journal is

The Royal Society of Chemistry 2012

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

acidity for the studied reaction along with the highly crystalline nature, higher BET surface area and pore volume of
SZ-650 (Table 1), which may play a signicant role by making
the reaction facile and active.
Under (B) reaction variables, decrease in conversion was
observed over both SZ-650 and SZ-700 catalysts (Table 4).
Though, the SZ-650 catalyst showed similar conversion to
SZ-600 under (A) reaction conditions, it reduced to 80%
under (B) reaction conditions. Because, its lower acidity and
Bronsted acid site concentration (compared to SZ-600) were
sucient to achieve 498% conversion when its concentration
was higher in reaction (A), however, with lower concentration
along with higher amount of methanol in reaction (B), the
number of acid sites available was insucient to interact with
the reactant molecules and thus to obtain complete conversion
of acid. The decreasing catalytic activities of SZ catalysts were
in good agreement with their decreasing Bronsted acid site
concentration as well as B/L ratio and total surface acidity
with an increase in calcination temperature (Table 1). These
results clearly revealed that the acidity and activity of a SZ
catalyst are strongly inuenced by the temperature at which it
has been calcined before undergoing the reaction along with
the reaction parameters. The pure ZrO2 sample, calcined at
400600 1C, showed only 2328% conversion of myristic acid
(Table 4) thus conrming the enhanced catalytic activity of SZ
catalysts.
2.2.4 Eect of alcohols. The reaction was carried out with
dierent short chain alcohols namely ethanol, n-propanol and
n-butanol over SZ-600 at (A) reaction parameters. Methanol
and ethanol having smaller carbon number resulted in similar
conversion (98%) at 60 1C after 7 h, however the conversion
was decreased successively with propanol (71%) and butanol
(69%) due to the inductive eect of increased carbon chain of
alcohol (Fig. 5).
A similar trend was found with caprylic acid26b and also has
been reported for SZ catalyzed transesterication of triglycerides10
and esterication of carboxylic acids of dierent chain lengths

Fig. 5 Esterication of myristic acid with dierent alcohols over the


SZ-600 catalyst. Reaction conditions: acid : methanol = 1 : 10; SZ =
0.5 wt%.

This journal is

The Royal Society of Chemistry 2012

with methanol using sulfuric acid and naonsilica composite33


suggesting that the lower reaction rates are due to steric
hindrance eects of the larger alkyl chains either of alcohols
or acids. However, in the present study the reduced conversion
of myristic acid with propanol and butanol was enhanced by
increasing the reaction temperature to 90 1C. The results
clearly envisage that methanol having a small carbon chain
is more economical for the esterication of fatty acid owing to
its lower boiling point that consumes lower thermal energy
and also in terms of comparatively inexpensive. The selectivity
for alkyl myristate was 100% irrespective of alcohol and
temperature. As excess of alcohol may lead to dehydration
or etherication in the presence of acid catalyst, the selectivity
of ester was further conrmed by checking the formation of
side products during the reaction of methanol with a SZ-600
catalyst at 60 1C for 510 h. No dehydrated or ether product
was detected by GC analysis under the conditions studied.
2.3 Re-usability of SZ catalysts
The used SZ-600 and SZ-650 catalysts were recovered from the
reaction mixture, washed with methanol, dried and activated
at 450 1C for 2 h before their re-use for further reaction cycle
under the similar reaction conditions (A).
The SZ-600 catalyst exhibited similar activity for myristic
acid esterication till ve reaction cycles without showing any
signicant decrease in the conversion (only a slight decrease to
94%). However, the activity of SZ-650 was decreased to 86%
after one re-cycle, which further decreased (71%) after ve
reaction cycles (Fig. 6). The sulfur content of the re-activated
SZ-600 and SZ-650 catalysts after ve reaction cycles was
decreased; 1.84 wt% and 0.604 wt%, respectively, compared
to fresh catalysts. However, IR spectra of re-activated catalysts were found to be similar to the fresh ones (Fig. 7).
Furthermore, the absence of any adsorbed organic species in
the IR spectra of re-activated SZ-600 and SZ-650 catalysts
indicated that the deactivation of the catalyst did not occur
due to the deposition of the product molecules on the active
acid sites of the catalyst.

Fig. 6 Re-usability of SZ-600 and SZ-650 catalysts after dierent runs.

Catal. Sci. Technol.

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

Fig. 7 IR spectra of (i) fresh and (ii) re-activated (after 5 cycles) (a)
SZ-600 and (b) SZ-650 catalysts.

We tried to nd out the possible reasons for decreased


activity of the SZ-650 catalyst as deactivation of a SZ catalyst
is supposed to be its major disadvantage though having strong
super acidity.
2.4

Deactivation of SZ catalysts

The successive decrease in the activity of SZ catalysts may be


due to (i) deactivation of active acid sites by water molecules
formed during the esterication reaction or/and (ii) leaching of
SO42 species from the catalyst in polar alcohol medium
during the course of reaction. Both possibilities have been
debated in the following sections.
2.4.1 Presence of water. We have checked the eect of
moisture present in methanol by repeating the reaction using a
fresh bottle of methanol of same purity. The initial conversion
of myristic acid was found to be signicantly higher (B1.6 times)
compared to the previous reaction studied under the similar
reaction conditions and B100% conversion was achieved
after 5 h instead of 7 h and was steady afterwards (Fig. 4c).
The results indicated the sensitivity of the esterication reaction of myristic acid towards the atmospheric moisture present
in alcohol. For rest of the studies, the fresh methanol was used
with precautions.
The eect of water was also studied by adding a small
amount of water (2000 ppm) to the reaction mixture. The
reaction showed lower conversion (9091%) of myristic acid
after 57 h (Fig. 8). Water has higher anity to interact
with the active acid sites and thus lowers the interaction of
methanol with active sites resulting in a decrease in the
conversion of acid. A similar retarding eect of water has
been observed in homogeneous acid catalyzed esterication of
FAs and also base catalyzed transesterication of triglycerides,
however, acid catalyzed esterication is more tolerant to water
compared to the latter.34 Kiss et al.35 have successfully used a
reactive distillation process to remove the water, however, the
reaction was done at higher temperature (130160 1C).
The results indicated the sensitivity of the esterication reaction of FAs towards the water molecules present in the
reaction system.
Catal. Sci. Technol.

Fig. 8 Eect of addition of water on the conversion of myristic acid


with methanol. Reaction conditions: acid : methanol = 1 : 10; SZ =
0.5 wt%

2.4.2 Leaching test. The leaching of SO42 species of SZ


catalysts especially in aqueous/alcoholic medium is a subject
of controversy, as some studies reported the hydrolysis of SZ
in pure aqueous medium36 but deactivation of SZ catalysts did
not occur by leaching of SO42 groups in the presence of small
amount of water in organic phase,9,36 whereas other reports
claimed the leaching of active SO42 groups from the catalyst
surface and transformed into H2SO4/HSO4 /SO42 ions in the
presence of alcohol10,11 or water.37
We have studied the leaching of SO42 ions of the catalyst
by employing two methods. (i) In one experiment, it was tested
during the esterication reaction. After 10 min of the reaction,
the catalyst was removed from the reaction mixture; and the
reaction was continued with the remaining solution without
having any catalyst. The plot of concentration of myristic acid
with time (Fig. 9) clearly showed an insignicant decrease in
concentration of myristic acid after removal of the catalyst as

Fig. 9 Concentration of myristic acid (%) without catalyst,


(J) catalyst removed after 10 min and (m) with catalyst. Reaction
conditions: acid : methanol = 1 : 10; SZ = 0.5 wt%

This journal is

The Royal Society of Chemistry 2012

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

compared to a linear decrease in myristic acid concentration in


the presence of catalyst. The autocatalysis (a blank test in the
absence of catalyst), under the similar reaction conditions,
showed nil conversion till 45 min and increased from 2.99%
(at 60 min) to 4.87% after 120 min (19% after 7 h, not shown
in gure). The initial rate of autocatalysis was signicantly
lower (0.5 mmol min 1) as compared to the catalytic reaction
(12.9 mmol min 1). The rate after removal of the catalyst
(2.4 mmol min 1) was observed to be ve times lower than
the reaction done in the presence of catalyst. These results
showed that no signicant leaching of SO42 ions of SZ
catalysts occur under the experimental conditions studied.
(ii) In a second experiment, a fresh SZ-600 catalyst was
stirred in methanol and distilled water separately (25 ml of
each) and pH was measured (Toshniwal, CL54) during 24 h.
The pH of the reaction mixture was found to decrease
signicantly after 1 h of stirring both in methanol and water,
however, it was not much aected further till 24 h (Fig. S6,
ESIw). Initially, the pH of the reaction mixture was higher in
water (4.3) compared to methanol (2.7), however, the decrease
in pH was observed in a similar range (38.4% in water and
38.6% in methanol) after 1 h of stirring.
It appeared from the decrease in pH by treatment with
water or methanol that the sulfate groups may be hydrolyzed
in the presence of OH groups on the catalyst surface9 or the
sulfate groups may be leached out during the initial rst
hour,10 which may not be strongly bonded with the zirconia
surface; however strongly bonded sulfate groups did not leach
out till further 24 h of stirring. To further conrm this, the
SZ-700 catalyst having less sulfur content and highly crystalline nature was checked for the same study, which also showed
a decrease in pH from 4.74 to 3.63 during the rst hour of
stirring in water. The pure ZrO2 sample having no sulfate
species also showed a decrease in pH from 5.76 to 4.885.04
during 12 hours of stirring in water, though it was lower
compared to SZ. Therefore, the decrease in pH (as also
reported by others) does not seem to be purely due to the
leaching of sulfate species; it may be due to the hydrolysis of
sulfate species and also due to the acidic metal oxide matrix.

2.5 Reaction rate


The kinetic prole of esterication of myristic acid with fresh
methanol (Fig. 10a) showed the continuous decrease in
concentration of myristic acid with time. The myristic acid
concentration [CA] was calculated by subtracting the consumed concentration from the initial concentration of myristic
acid. A fast linear decrease in myristic acid concentration was
observed till 60 min. The reaction rate (n) was determined by
applying linear analysis to the plot of decreasing [CA] with
time by the equation, n = d[CA]/dt. The initial rate was
12.9 mmol min 1. Turnover frequency (TOF, moles of fatty
acid converted per mole of acid site concentration per hour) at
100% conversion was 4.44 per hour.
The kinetic experiments with o10 min reaction time were
performed to have a view on apparent induction period of the
reaction. The kinetic prole in Fig. 10b shows the conversion
of myristic acid within 1 min of the reaction, a fast decrease in
[CA] between 1 and 2 min, which extended to 4 min and after
that it became slow. It is to be noted that while performing the
reaction, the catalyst was added after mixing the acid and
methanol by stirring at ambient temperature. This time was
taken as zero time. After that the reaction mixture was heated
at the desired temperature (60 1C), which was already maintained at 60 1C (to avoid the shooting of temperature (B45 1C)
beyond 60 1C that normally occurs during the heating of the
system). This was kept in mind because of low boiling point of
methanol (6465 1C). Thus, the apparent induction period of the
studied reaction is just the heating of the system; it will be slightly
extended during the heating of the system to the desired temperature. The initial reaction rate of this period was determined
to be signicantly higher (38.8 mmol min 1).
2.6 Comparison of catalytic performance with other acids
Based on literature reports9,1113,36 we have focused on SZ in the
present study. However, we have also carried out the reaction
with concentrated H2SO4 and heterogeneous ion exchange resin
Amberlyst-15 to have a comparable view on the catalytic activity.
The results (Fig. S7, ESIw) showed the activity in the order of

Fig. 10 Esterication of myristic acid with fresh methanol (a) conversion/concentration of myristic acid with time (b) apparent induction period.
Reaction conditions: acid : methanol = 1 : 10; SZ = 0.5 wt%

This journal is

The Royal Society of Chemistry 2012

Catal. Sci. Technol.

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

View Online

H2SO4 4 SZ 4 Amberlyst-15. Concentrated H2SO4, used as a


benchmark catalyst, is not eco-friendly and cannot be readily
re-cycled. To overcome the disadvantages associated with
liquid acid catalysts, heterogeneous catalysts are the sustainable alternatives, though having lower catalytic performance.
The initial rate of esterication reaction was slow with SZ
catalysts (12.9 mmol min 1) as compared to conc. H2SO4
(1415 mmol min 1), the nal conversion after 2 h was in a
comparable range. These results clearly showed the potential
of SZ solid acid catalyst near to the benchmark catalyst.
Amberlyst-15 showed signicantly lower activity (56 mmol min 1)
under the reaction conditions studied and was comparable with
esterication of oleic acid with methanol at 80 1C.38
A comparison of catalytic activity of SZ used in the present
study with reported studies of SZ or other solid acid catalysts
gave fairly good performance (716 mmol min 1 g 1) as compared to reported SZ (125 mmol min 1 g 1) and SZ incorporated on SBA-15 (175 mmol min 1 g 1) for methyl palmate
and laurate formation at 68 1C.39 TOF of the present reaction
was also found to be higher (4.44 per hour) as compared to
SZ (1.92.5 per hour) and SZ incorporated on SBA-15
(2.63.4 per hour) for methyl palmate and laurate (calculated
from data given in ref. 39). Similarly, the activity was observed
to be higher (716 mmol min 1 g 1) as compared to other SZ
(150 mmol min 1 g 1) and sugar catalyst (478 mmol min 1 g 1)
for methyl oleate formation at 80 1C.38 The conversion rates
were also observed in a similar range for esterication of lauric
acid with 2-ethyl hexanol by the SZ (SSZr-1) catalyst, studied
at 170 1C with 10 wt% of catalyst.8 The authors reported
similar results for esterication of lauric and myristic acids
with iso-propanol. TOF values of the reported reactions may
be higher. A high reaction rate for esterication of myristic
acid with methanol was observed over SZ18 (7585% conversion in 5 min) at 22 bar in a Parr reactor at 120170 1C. The
high temperature and pressure are important factors to signicantly enhance the reaction rates besides the catalyst properties.
Therefore, a proper comparison of catalytic activity, though
normalized with acid site concentration, can only be viable
when the reactions have been performed at comparable reaction
parameters such as temperature, catalyst amount and acid to
alcohol ratio9,18,40,41 along with the carbon chain length of
alcohol8,20,41 and carboxylic33/fatty acid.19,20
2.7

Structureactivity co-relation

The co-relation between acidic features in terms of total


surface acidity, Bronsted acidity, Bronsted acid site concentration, B/L ratio and the activity of all three SZ catalysts
reveals that esterication of myristic acid with short chain
alcohols is mainly Bronsted acid catalyzed reaction. The
activity decreases with a decrease in Bronsted acidity in either
terms. Furthermore, a SZ catalyst with lower number of weak
and moderate acid sites is also able to demonstrate the
esterication activity similar to a SZ catalyst having comparatively higher number of strong acid sites; however, the latter is
requisite to retain the original activity of re-used SZ catalysts
along with a higher B/L ratio. The dierence in the activity of
re-used catalysts clearly revealed that re-usability of SZ catalysts strongly depends on the acidity of the catalysts.
Catal. Sci. Technol.

3. Experimental
3.1 Materials
Zirconium n-propoxide (70 wt% in n-propanol) was procured
from Sigma-Aldrich, Germany; n-propanol (99.5%), n-butanol
(99.5%), aqueous ammonia solution (25%) and concentrated
sulfuric acid (98%) were procured from s.d. Fine chemicals,
India. Methanol (99%) was purchased from Ranbaxy, India
and Merck, India. Myristic acid (99.5%) and ethanol (99.9%)
were procured from Spectrochem Pvt. Ltd., India.
3.2 Catalyst synthesis and characterization
A sulfated zirconia catalyst was prepared using a two-step
sol-gel technique as described previously.26a In a typical
synthesis procedure, Zr(OPr)4 was hydrolyzed by aqueous
ammonia and the obtained gel was treated with H2SO4 after
oven drying at 120 1C for 12 h. The sulfated powder was
calcined at 600700 1C for 4 h and designated SZ-T, where T
represents the calcination temperature.
PXRD patterns were obtained using a Philips Xpert
diractometer. The crystallite size was determined from the
characteristic peak of tetragonal phase (2y = 30.22) using the
Scherrer equation,42 crystallite size = Kl/W cos y, where K is
the Scherrer constant (0.9), l = 1.5406 A (CuKa radiation),
W = Wb
Ws; Wb is the broadened prole width of the
experimental sample and Ws is the standard prole width of
the reference silicon sample and y is the angle of diraction.
TEM micrographs were obtained using a JEOL JEM 2100
transmission electron microscope by dispersing the catalyst
sample in ethanol by sonication and deposited on a Cu grid
coated with carbon lm. IR spectra were recorded on a Perkin
Elmer GX spectrophotometer. The bulk sulfur present before
and after calcination at dierent temperatures was analyzed by
ICP emission spectroscopy using Perkin-Elmer, Optima 2000
DV. BET surface area, pore volume and pore diameter were
calculated from nitrogen sorption isotherms after pre-activation
at 120 1C for 4 h at 196 1C on ASAP 2010, Micromeritics.
NH3-TPD was used to estimate the total surface acidity, i.e.
strength and number of acid sites present in the catalysts using
Micromeritics Pulse Chemisorb 2720 as described earlier.26a
Vapor phase cyclohexanol dehydration in a xed bed reactor
was used as a model reaction to assess the Bronsted acidity of
the catalysts.26b Diuse Reectance FT-IR spectroscopy
(DRIFT) was used to dierentiate between Bronsted and
Lewis acidity of pyridine adsorbed SZ samples using Perkin
Elmer GX equipped with DRIFT Selector accessory (Graseby
Specac, P/N 19900 series).25e The spectra were recorded at
room temperature (B27 1C) to 450 1C after holding at each
temperature for 10 min, thus allowing sucient time for
pyridine desorption.
3.3 Catalytic activity
Esterication of myristic acid was carried out in a liquid phase
batch reactor.26b In a typical reaction procedure, required
amounts of acid, alcohol, and catalyst were taken in a round
bottom ask and the suspension was magnetically stirred
(600 rpm) in an oil bath maintained at constant temperature
(1 1C) in the temperature range of 4060 1C. The reactions
This journal is

The Royal Society of Chemistry 2012

View Online

were performed in duplicate with 2% variations in conversion. The samples were withdrawn at regular intervals and the
reaction product was analyzed using a gas chromatograph (HP
6890) having a DB-225 capillary column (20 m length, 100 mm
diameter and 0.10 mm lm thickness) and a FID detector.

Downloaded on 01 November 2012


Published on 31 July 2012 on http://pubs.rsc.org | doi:10.1039/C2CY20462B

4. Conclusions
The acidity and activity of a SZ catalyst for esterication of
myristic acid with methanol and other short chain alcohols are
strongly inuenced by its calcination temperature along with
the reaction parameters mainly acid to alcohol ratio, catalyst
concentration and reaction temperature. Among the structural,
textural and acidic properties of the catalyst, the acidic
features play a major role in its activity and re-usability
behaviour for the studied reaction.
The present study reveals a very remarkable result that the
SZ-catalyst having higher number of Bronsted acid sites could
be re-used for ve reaction cycles without signicant loss in
activity for myristic acid esterication, whereas the SZ-catalyst
having less number of acid sites showed decrease in activity
after ve reaction cycles. This explicitly conrms the requirement of more number of Bronsted acid sites for better
re-usability of SZ catalysts for esterication of myristic acid.
The present study concludes that the requirement of a very
small concentration of the catalyst yielding maximum conversion and selective formation of the ester at lower temperature
within reasonable reaction time makes SZ an appealing
catalyst for the synthesis of FAAEs. In addition, eco-friendly,
recyclable SZ catalysts may nd wide applications in reactions
where conc. H2SO4 is currently being used as a catalyst.

Acknowledgements
The authors are thankful to CSIR Network Programme on
Inorganic Materials for Diverse Application and to Analytical
Science discipline for catalyst characterization analysis. The
authors are also thankful to Dr Ram S. Shukla for helping in
initial reaction rate determination.

Notes and references


1 J. Beare-Rogers, A. Dieenbacher and J. V. Holm, Lexicon of lipid
nutrition (IUPAC Technical Report), Pure Appl. Chem., 2001, 73,
685744.
2 T. A. Hughes, M. Heimberg, X. Wang, H. Wilcox, S. M. Hughes,
E. A. Tolley, D. M. Desiderio and J. T. Dalton, Metabolism, 1996,
45, 11081118.
3 R. W. Johnson and E. Fritz, in Fatty Acids in Industry: Processes,
Properties, Derivatives, Applications, ed. Marcel Dekker, New
York, 1988.
4 Commonwealth of Australia Gazette no. APVMA 37, March
2006, p. 13.
5 S. Hartwig, V. Auwarter and F. Pragst, Alcohol Alcohol., 2003, 38,
163167.
6 E. Lotero, Y. Liu, D. E. Lopez, K. Suwannakarn, D. A. Bruce and
J. G. Goodwin Jr., Ind. Eng. Chem. Res., 2005, 44, 53535363.
7 J. A. Melero, J. Iglesias and G. Morales, Green Chem., 2009, 111,
2851308.
8 M. L. Grecea, A. C. Dimian, S. Tanase, V. Subbiah and
G. Rothenberg, Catal. Sci. Technol., 2012, 2, 15001506.
9 A. A. Kiss, A. C. Dimian and G. Rothenberg, Adv. Synth. Catal.,
2006, 348, 7581.

This journal is

The Royal Society of Chemistry 2012

10 K. Suwannakarn, E. Lotero, J. G. Goodwin Jr. and C. Lu,


J. Catal., 2008, 255, 279286.
11 D. E. Lopez, J. G. Goodwin Jr., D. A. Bruce and S. Furuta, Appl.
Catal., A, 2008, 339, 7683.
12 (a) S. Furuta, H. Matsuhashi and K. Arata, Biomass Bioenergy,
2006, 30, 870873; (b) S. Furuta, H. Matsuhashi and K. Arata,
Catal. Commun., 2004, 5, 721723.
13 K. F. Yee, J. C. S. Wu and K. T. Lee, Biomass Bioenergy, 2011, 35,
17391746.
14 M. Hino, S. Kobayashi and K. Arata, J. Am. Chem. Soc., 1979,
101, 64396441.
15 B. M. Reddy and M. K. Patil, Chem. Rev., 2009, 109, 21852208.
16 G. D. Yadav and J. J. Nair, Microporous Mesoporous Mater.,
1999, 33, 148.
17 X. Song and A. Sayari, Catal. Rev. Sci. Eng., 1996, 38, 329412.
18 D. Rattanaphra, A. P. Harvey, A. Thanapimmetha and
P. Srinophakun, Renewable Energy, 2011, 36, 26792686.
19 X. Hu, Z. Zhou, D. Sun, Y. Wang and Z. Zhang, Catal. Lett.,
2009, 133, 9096.
20 K. Srilatha, N. Lingaiah, P. S. S. Prasad, B. L. A. P. Devi, R. B.
N. Prasad and S. Venkateswar, Ind. Eng. Chem. Res., 2009, 48,
1081610819.
21 J. C. Juan, J. Zhang and M. A. Yarmo, Catal. Lett., 2008, 126,
319324.
22 T. Yalcinyuva, H. Deligoz, I. Boz and M. A. Gurkaynak, Int. J.
Chem. Kinet., 2008, 40, 136144.
23 A. Sakthivel, K. Komura and Y. Sugi, Ind. Eng. Chem. Res., 2008,
47, 25382544.
24 K. Mantri, K. Komura and Y. Sugi, GreenChem., 2005, 7,
677682.
25 (a) R. V. Jasra, B. Tyagi and M. K. Mishra, US Pat., 2009 7, 544
831 B2, Aug 2009; (b) B. Tyagi, M. K. Mishra and R. V. Jasra,
J. Mol. Catal. A: Chem., 2010, 317, 4145; (c) B. Tyagi,
M. K. Mishra and R. V. Jasra, J. Mol. Catal. A: Chem., 2009,
301, 6778; (d) B. Tyagi, M. K. Mishra and R. V. Jasra, Catal.
Commun., 2006, 7, 5257; (e) B. Tyagi, M. K. Mishra and
R. V. Jasra, J. Mol. Catal. A: Chem., 2008, 286, 4146;
(f) B. Tyagi, M. K. Mishra and R. V. Jasra, J. Mol. Catal. A:
Chem., 2007, 276, 4756.
26 (a) K. Saravanan, B. Tyagi and H. C. Bajaj, J. Sol-Gel Sci.
Technol., 2012, 61, 275280; (b) K. Saravanan, B. Tyagi and
H. C. Bajaj, J. Sol-Gel Sci. Technol., 2012, 62, 1317.
27 S. J. Gregg and K. S. W. Sing, Adsorption, Surface Area and
Porosity, Academic Press, New York, 2nd edn, 1982.
28 T. Yamaguchi, T. Jin and K. Tanabe, J. Phys. Chem., 1986, 90,
31483152.
29 B. T. Loveless, A. Gyanani and D. S. Muggli, Appl. Catal., B,
2008, 84, 591597.
30 A. L. Cardoso, R. Augusti and M. J. D. Silva, J. Am. Oil Chem.
Soc., 2008, 85, 555560.
31 M. Hino, S. Takasaki, S. Furuta, H. Matsuhashi and K. Arata,
Catal. Commun., 2006, 7, 162165.
32 S. Ramu, N. Lingaiaha, B. L. A. P. Devi, R. B. N. Prasad,
I. Suryanarayanaa and P. S. S. Prasad, Appl. Catal., A, 2004,
276, 163168.
33 Y. Liu, E. Lotero and J. G. Goodwin Jr., J. Catal., 2006, 243,
221228.
34 D. Kusdiana and S. Saka, Bioresour. Technol., 2004, 91, 289295.
35 A. A. Kiss, A. C. Dimian and G. Rothenberg, Energy Fuels, 2008,
22, 598604.
36 F. Omata, A. C. Dimian and A. Bliek, Chem. Eng. Sci., 2003, 58,
31753185.
37 X. B. Li, K. Nagaoka and J. A. Lercher, J. Catal., 2004, 227,
130137.
38 M. H. Zong, Z. Q. Duan, W. Y. Lou, T. J. Smith and H. Wu,
Green Chem., 2007, 9, 434437.
39 X. R. Chen, Y. H. Ju and C. Y. Mou, J. Phys. Chem. C, 2007, 111,
1873118737.
40 A. A. Kiss, F. Omota, A. C. Dimian and G. Rothenberg, Top.
Catal., 2006, 40, 141150.
41 M. C. de Jong, R. Feijt, E. Zondervan, T. A. Nijhuis and A. B. de
Haan, Appl. Catal., A, 2009, 365, 141147.
42 B. D. Cullity and S. R. Stock, Elements of X-ray Diraction,
Prentice Hall: Upper Saddle River, NJ, 3rd edn, 2001, p. 388.

Catal. Sci. Technol.

Das könnte Ihnen auch gefallen