ArticlePDF Available

Biodiesel production using heterogeneous catalysts

Authors:
  • Research & Development Centre, Indian oil corporation, Faridabad
  • indian oil corporation

Abstract and Figures

The production and use of biodiesel has seen a quantum jump in the recent past due to benefits associated with its ability to mitigate greenhouse gas (GHG). There are large number of commercial plants producing biodiesel by transesterification of vegetable oils and fats based on base catalyzed (caustic) homogeneous transesterification of oils. However, homogeneous process needs steps of glycerol separation, washings, very stringent and extremely low limits of Na, K, glycerides and moisture limits in biodiesel. Heterogeneous catalyzed production of biodiesel has emerged as a preferred route as it is environmentally benign needs no water washing and product separation is much easier. The present report is review of the progress made in development of heterogeneous catalysts suitable for biodiesel production. This review shall help in selection of suitable catalysts and the optimum conditions for biodiesel production.
Content may be subject to copyright.
Review
Biodiesel production using heterogeneous catalysts
Surbhi Semwal
a
, Ajay K. Arora
b
, Rajendra P. Badoni
a
, Deepak K. Tuli
b,
a
College of Engineering, University of Petroleum & Energy Studies, Dehradun 248007, India
b
Research & Development Centre, Indian Oil Corporation Limited, Sector-13, Faridabad 121007, India
article info
Article history:
Received 5 May 2010
Received in revised form 21 September
2010
Accepted 18 October 2010
Available online 23 October 2010
Keywords:
Biodiesel
Homogeneous catalyst
Heterogeneous catalyst
abstract
The production and use of biodiesel has seen a quantum jump in the recent past due to benefits associ-
ated with its ability to mitigate greenhouse gas (GHG). There are large number of commercial plants pro-
ducing biodiesel by transesterification of vegetable oils and fats based on base catalyzed (caustic)
homogeneous transesterification of oils. However, homogeneous process needs steps of glycerol separa-
tion, washings, very stringent and extremely low limits of Na, K, glycerides and moisture limits in biodie-
sel. Heterogeneous catalyzed production of biodiesel has emerged as a preferred route as it is
environmentally benign needs no water washing and product separation is much easier. The present
report is review of the progress made in development of heterogeneous catalysts suitable for biodiesel
production. This review shall help in selection of suitable catalysts and the optimum conditions for bio-
diesel production.
Ó2010 Elsevier Ltd. All rights reserved.
1. Introduction
Bio-based alternative fuels such as ethanol, biodiesel have been
in focus for the reasons which are by now well understood. Heavy
consumption of fossil resources, effect on global warming and con-
cerns of energy security are main drivers for growth of biofuels. Re-
cent studies on life cycle analysis (LCA) of biodiesel have shown a
very appreciable reduction of greenhouse gas (GHG) by their use as
a blend component of transport fuel. Biodiesel produced by transe-
sterification of vegetable oils and animal fats using homogeneous
base catalyst (Fig. 1) has seen several folds increase in last few
years for their commercial production and use as a blending com-
ponent in transport fuels.
Fatty acid methyl esters (FAME) have found favour for use as a
blend component of petro-diesel fuel due to lack of aromatics, neg-
ligible sulfur content, higher lubricity and very high cetane values
(Dorado et al., 2003). FAME (Biodiesel) mixes freely in all propor-
tions with petro-diesel and its use has been approved by almost
all the major automotive manufactures. Biodiesel can be used in
conventional compression ignition engines, which need almost
no modification. Though biodiesel has been approved for use in
automotives as a blend with normal petro-diesel, there are very
stringent quality norms prescribed by several countries. ASTM
specifications listed and detailed by Sarin et al. (2007), which
any biodiesel must meet before it can be used as an auto fuel com-
ponent. There are very low limits on Na/K, organic/inorganic acids,
phosphorous, glycerides and water content. Therefore, biodiesel
production processes need to have in-built capability to meet these
specifications.
The most widely used industrial method for the commercial
production of biodiesel from vegetable oils/fats is a base catalyzed
transesterification process using KOH or NaOH as the homoge-
neous catalyst and MeOH as the lower alcohol (Fig. 1). The advan-
tage of this process is production of methyl esters at very high
yields under mild conditions and reaction generally takes about
an hour for completion (Meher et al., 2006). Several oils, both edi-
ble and non-edible such as sunflower oil (Arzamendi et al., 2008),
palm (Li and Xie, 2006) and jatropha (Tiwari et al., 2007) have been
transesterified for biodiesel production. However, major quality re-
lated problems were encountered and it was main hindrance for
large scale industrial production of biodiesel by homogeneously
catalyzed transesterification. Production costs were rather high
(Ma and Hanna, 1999) as the process involved number of washing
and purification steps in order to meet the stipulated quality. It
was quite difficult to remove the traces the K/Na remaining in
the product and separation of glycerin also posed technical chal-
lenges. The higher amount of water used in washing and conse-
quent treatment of the resulting effluent added to the overall
process cost. Any commercial biodiesel plant must have the in-
built capability to handle a variety of different feedstocks which
may differ very widely in quality. The vegetable oils may be from
edible sources, non-edible sources, waste cooking oils, animal fats,
algae, fungi etc. (Canakci, 2007; Granados et al., 2007; Ji et al.,
2006; Karmee and Chadha, 2005).
In Europe and US, the primary sources for producing biodiesel
are edible oils like rapeseed, sunflower, and soybean. In countries
0960-8524/$ - see front matter Ó2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2010.10.080
Corresponding author. Tel.: +91 129 22 94 273; fax: +91 129 22 85 340.
E-mail address: tulidk@iocl.co.in (D.K. Tuli).
Bioresource Technology 102 (2011) 2151–2161
Contents lists available at ScienceDirect
Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech
like India, non-edible oils like jatropha and karanjia are being pro-
moted on a very large scale, as these can be grown on marginal and
waste lands (Azam et al., 2005). Several other non-edible seeds like
Guizotia abyssinica (Sarin et al., 2009a) have also been evaluated
for their biodiesel potential. The conventional biodiesel production
process of base catalyzed homogeneous transesterification face dif-
ficulties to handle multiple feed stocks. Oils (nonedible) with high-
er fatty acid content lead to formation of soap and consequent loss
of oil and problems of product separation (Kwiecien et al., 2009).
Due to these issues a large number of alternative methods were
developed. These include supercritical process (Minami and Saka,
2006) and enzymatic process (Shimada et al., 2002). Supercritical
process is also one of the promising methods for biodiesel produc-
tion as this process is very fast and is carried out without catalyst.
Some production plants in Europe use this technology, but due to
high temperature and pressure requirement of this process, it
translates to higher capital costs and that restricts its commercial
utilization. Sharma et al. (2006) explored a single pot process for
transesterification of jatropha oil.
Enzyme based transesterification is also one of the option for
biodiesel production and is generally carried out at moderate tem-
perature with high yields. Lipase enzymes (used with different
supports by immobilization or encapsulation etc.) are used for
transesterification reaction (Caballero et al., 2009; Macario et al.,
2007, 2009). This process can tolerate free fatty acid and water
without soap formation and thereby making separation of biodie-
sel and glycerol easier. Enzyme cost and its deactivation due to
feed impurities are major hindrance for commercial viability of this
process (Dizge et al., 2009).
Biodiesel synthesis using solid catalysts instead of homoge-
neous liquid catalyst could potentially lead to economical produc-
tion costs because of reuse of the catalyst (Suppes et al., 2004) and
offer the possibility for carrying out both transesterification and
esterification simultaneously (Furuta et al., 2004).
Additional benefit with solid based catalyst is the lesser con-
sumption of catalyst. As per studies, for production of 8000 tonnes
of biodiesel, 88 tones of sodium hydroxide may be required
(Mbaraka and Shanks, 2006), while only 5.7 tonnes of solid
supported MgO is sufficient for production of 100,000 tonnes of
biodiesel (Dossin et al., 2006).
One disadvantage with use of solid catalyst is the formation of
three phases together with oil and alcohol, which leads to diffusion
limitations thus decreasing the rate of the reaction (Mbaraka and
Shanks, 2006). This mass transfer difficulty is overcome by using
a co-solvent such as tetrahydrofuran (THF), dimethyl sulfoxide
(DMSO), n-hexane and ethanol, which assist miscibility of oil and
methanol leading to increase in the rate of reaction. Use of catalyst
supports, which can provide more specific surface area and pores
for active species, where they can anchor and react with large tri-
glyceride molecules is another solution for encountering the poor
mass transfer (Zabeti et al., 2009). Some researchers (Di Serio
et al., 2008) have reviewed the significance of solid catalysts for
biodiesel production. However, till date few detailed kinetic stud-
ies and the mechanism of acid and base catalysts using solid cata-
lysts have been reported in the literature (Furuta et al., 2004;
Suppes et al., 2004).
Chemistry of heterogeneous catalyst reported, includes metal
hydroxides (Dalai et al., 2006), metal complexes (Abreu et al.,
2003), metal oxides such as calcium oxide (Granados et al.,
2007), magnesium oxide (Wang and Yang, 2007), zirconium oxide
(Jitputti et al., 2006), zeolites, hydrotalcites and supported cata-
lysts (Xie and Huang, 2006). These types of catalysts have been
investigated as solid catalysts which overcome some of the draw-
back on use of homogeneous catalysts. The order of activity among
alkaline earth oxide catalysts was observed to be BaO > SrO > -
CaO > MgO (Cantrell et al., 2005).
The present review discusses the use of acid, base, acid–base so-
lid catalysts such as metal oxides, supported catalysts and zeolites
etc., and enzymatic catalysts for biodiesel synthesis. This review
paper presents a comparative description of continuous biodiesel
production processes, through transesterification reaction using
acid, base, acid–base heterogeneous catalysts and enzymatic cata-
lysts so that proper catalyst and optimum reaction conditions can
be selected.
2. Heterogeneous catalysts
2.1. Basic solid catalysts
Various basic metal oxide type catalysts have been reported in
literature for biodiesel synthesis. Some of the high performing cat-
alyst preparations and their application in biodiesel synthesis are
summarized here.
Liu et al. (2007) studied SrO metal oxide for transesterification
of soybean oil. Catalyst preparation was carried out by calcinations
of SrCO
3
at 1200 °C for 5 h. SrO has strong basicity H
= 26.5 and
have BET surface area of 1.05 m
2
/g. The conversion obtained was
95% at temperature of 65 °C, catalyst content of 3 wt.%, molar ratio
of methanol to oil of 12:1 and reaction time of 30 min. Further, bio-
diesel yield was only slightly reduced when the SrO catalyst is sub-
sequently reused for 10 cycles. Mechanism given by authors is as
described in Fig. 2. The main step is formation of ionic complex
by SrO with methanol.
Liu et al. (2008) studied transesterification of soybean oil to bio-
diesel using CaO as a solid catalyst. The BET surface area of the cat-
alyst was 0.56 m
2
/g. The reaction was carried out using 12:1 M
ratio of methanol to oil, 8 wt.% catalyst concentration at 65 °C. Bio-
diesel yield (95%) was obtained when reaction was carried out for
3 h. The authors also reported comparative activity of CaO with
K
2
CO
3
/
c
Al
2
O
3
and KF/
c
Al
2
O
3
catalysts. Preparation of these cata-
lysts was carried out by an impregnation method with the help
of aqueous solution of potassium carbonate/potassium fluoride
and then calcination of impregnated catalysts at 550 °C for 5 h. It
was observed that CaO maintained sustained activity for longer
time (20 cycles) after repeated use and biodiesel yield was also
not affected, while K
2
CO
3
/
c
Al
2
O
3
and KF/
c
Al
2
O
3
catalysts were
not able to maintain activity and biodiesel yield also got affected
after every use. This was because that the alkali metal compounds
dissolved in methanol, which reduced the active ingredients and
thereby decreasing biodiesel yield in the subsequent experiments.
It was also observed in this study that the presence of water, if in
small amount of about 2.8 by wt.% of soybean oil, act as promoter,
but if amount of water increases (more than 2.8 by wt.% of soybean
oil) it hydrolyzed FAME under basic conditions and also induced
soap formation.
The catalytic activity of activated calcium oxide was also evalu-
ated by Granados et al. (2007) for production of biodiesel by
H
2
C
HC
OCOR'
OCOR''
H
2
COCOR'''
3 ROH
H
2
C
HC
OH
OH
H
2
COH
ROCOR'
ROCOR''
ROCOR'''
catalyst
triglyceride alcohol mixture of
alkyl esters
glycerol
R', R'', R''' = Hydrocarbon chain ranging from 15 to 21 carbon atoms
(FAME)
Fig. 1. Transesterification of vegetable oil.
2152 S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161
transesterification of sunflower oil in batch reactor at 13:1 metha-
nol to oil molar ratio, 3 wt.% catalyst content at 60 °C. Under these
conditions, reaction was complete in 100 min giving 94% conver-
sion. The specific surface area of catalyst was 32 m
2
/g and mean
pore size diameter (MPS) was approx. 25–30 nm. The authors ob-
served poisoning of active surface site of CaO by the atmospheric
H
2
O and CO
2
. Therefore, to improve catalytic activity of CaO, it
was subjected to an activation treatment at high temperature
(P700 °C) before the reaction and as a result of this, the main poi-
soning species (the carbonate group) from the surface was re-
moved. When catalyst was activated at high temperature, some
leaching of the active species was observed. However, leaching
amount did not result in significant reduction of catalyst activity
and the catalyst was reusable for 8 cycles. However, yield of FAME
reduced from more than 90% in the first cycle to 80% in the second
cycle and thereafter the performance stabilized.
Veljkovic et al. (2009) described the kinetics of CaO heteroge-
neously catalyzed methanolysis of sunflower oil. The optimal
CaO calcination temperature was 550 °C. They observed 98% yield
in the transesterification with 6:1 M ratio of sunflower oil to meth-
anol, 1 wt.% catalyst (based on oil wt.) at 60 °C within 2 h reaction
time.
Sarin et al. (2009b) reported use of seashell and eggshells as a
catalyst for production of biodiesel from various feed stocks such
as jatropha, castor, sunflower, soybean, rapeseed, cotton, corn,
coconut oils etc., in a batch and continuous reactor. The catalyst
combination of seashell and eggshells contained between 10–90%
and 90–10% respectively. The reaction was performed using
1 mol of vegetable oil and 6 mol of methanol and 4 wt.% of catalyst
composition. 98% conversion was achieved within 2 h.
Kawashima et al. (2009) studied catalytic activity of calcium
oxide (CaO) as a heterogeneous catalyst for biodiesel production
by the transesterification of rapeseed oil. The author pretreated
CaO with methanol for activation. CaO was activated with metha-
nol at 25 °C for 1.5 h so that small amount of CaO could be con-
verted into Ca(OCH
3
)
2
, which exhibits a higher catalytic activity
than non-activated CaO. Rapeseed oil was thus transesterified
using Ca(OCH
3
)
2
to produce FAME and glycerine. During the
transesterification reaction, the produced glycerin reacted with
CaO at 60 °C, and a CaO-glycerin complex was formed as secondary
catalyst, which then accelerated the transesterification reaction.
While this CaO-glycerin complex exhibited a high catalyst activity,
the reaction advanced further and generated more glycerin. To
determine the exact pattern of catalytic activity, XRD measure-
ments of activated CaO, non-activated CaO, Ca(OH)
2
, and
Ca(OCH
3
)
2
were performed and it was observed that XRD spectrum
of activated CaO was similar to that of non-activated CaO but
exhibiting small diffraction peak attributed to Ca(OCH
3
)
2
and
Ca(OH)
2
. This was responsible for the observed differences in the
catalytic activity and the basic strengths of non-activated CaO,
Ca(OH)
2
. The activated CaO has basic strength in the range of
10.1–11.1. While Ca(OCH
3
)
2
had a high basic strength in the range
of 11.1–15.0 and these results explained the reasons why
Ca(OCH
3
)
2
exhibited a higher catalytic activity for the transesteri-
fication reaction than CaO and Ca(OH)
2
.
Kouzu et al. (2008) studied CaO catalyst for transesterification
of soybean oil at 12:1 M ratio of methanol to oil at 500 rpm and
at reflux temperature for 2 h in glass batch reactor and achieved
93% biodiesel yield. CaO was obtained after calcination of pulver-
ized lime stone at 900 °C for 1.5 h. Calcium diglyceride and calcium
methoxide were used as reference samples. Further on comparison
it was observed that the BET surface area of fresh CaO is 13 m
2
/g
whereas surface area of CaO collected after conversion was
11 m
2
/g. While the BET surface area of reference samples such as
calcium diglyceroxide and calcium methoxide were 11.3 m
2
/g
and 44 m
2
/g, respectively.
Catalytic activity of calcium based metal oxides such as CaTiO
3
,
CaMnO
3
,Ca
2
Fe
2
O
5
, CaZrO
3
and CaO–CeO
2
in the methanolysis of
rapeseed oil was studied by Kawashima et al. (2008). The authors
also studied the change of activity on replacement of Ca with bar-
ium, magnesium, or lanthanum. The reaction was carried out in a
batch reactor at 60 °C with 6:1 M ratio of methanol to rapeseed
oil for 10 h, resulting in yield of 79–92%. It was found that CaZrO
3
and CaO–CeO
2
show high durability, ester yields greater than 80%
and has the potential to be used in biodiesel production processes
as heterogeneous base catalysts. For synthesis of CaTiO
3
, an equi-
molar mixture of TiO
2
and CaCO
3
was milled in an agate mortar
then mixture calcined in air to 500 °C and subsequently at
1050 °C for 2 h. For preparing Ca
2
Fe
2
O
5
,Fe
2
O
3
and CaCO
3
were
milled with molar ratio of 1:2 and calcined in air to 900 °C and then
CH
3
-OH
Sr OSr O
CH
3
O H
R
1
O
OR R
1
OCH
3
OR
OSr O
H
R
1
CH
3
O
ROH
OSr O
CH
3
OH CH
3
O
R
1
OCH
3
O
ROH
Sr O
CH
3
O H
R
1
OCH
3
OR
OSr O
H
R
1
OCH
3
OR
OR
1
CH
3
O
ROH
O
R
1
CH
3
O
ROH
O
Fig. 2. Mechanism of SrO catalyst transesterification.
S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161 2153
at 1050 °C for 4 h. Due to calcination step at high temperature, the
surface area of each of the catalysts was small and varied from
7.7 m
2
/g for MgCeO
3
to 0.71 m
2
/g for Ca
2
Fe
2
O
5
. The basic strengths
of CaTiO
3
were in the range of 6.8–7.2. CaMnO
3
,Ca
2
Fe
2
O
5
, CaZrO
3
,
and CaCeO
3
showed the highest basic strength, while Ba, Mg, and
La series catalysts had weaker basic strength. Hence, Ca series cat-
alysts exhibit the high catalytic activity for the transesterification
reaction.
MgO-catalyzed transesterification reaction, at industrially rele-
vant conditions was reported by Dossin et al. (2006) in batch and
continuous stirred tank reactors. A kinetic model based on the
three steps ‘Eley–Rideal’ type mechanism assuming methanol
adsorption as rate-determining step was proposed. Two processes
were simulated, first for transesterification of ethyl acetate with
methanol in a batch slurry reactor and second, transesterification
of triolein with methanol to form methyl oleate in a continuous
slurry reactor and results were used to simulate biodiesel produc-
tion from rapeseed oil. In a continuous stirred reactor volume of
25 m
3
containing 5700 kg of MgO catalyst continuous production
of 100,000 tonnes of biodiesel per year can be achieved. These re-
sults were compared by the author with homogeneously catalyzed
transesterification processes.
K
2
CO
3
supported on MgO catalyst was prepared by mixing
K
2
CO
3
and MgO as carrier in a mortar. The mixture calcined at
600 °C for 3 h, thus forming catalyst for synthesis of biodiesel from
soybean oil with the yield of 99.5% (Liang et al., 2009). These re-
sults indicate that carriers increased the reaction yield and basic
carriers have higher activities than acidic carriers. The catalytic
activity of the K
2
CO
3
/MgO was higher than that of K
2
CO
3
due to
the interaction between K
2
CO
3
and MgO and because of the high
degree of dispersion of the active sites on the surface of MgO.
The maximum activity of catalyst was obtained when the loading
ratio was 0.7 and after 2 h reaction time the maximum conversion
was achieved in transesterification when operating parameters are
set at 70 °C, 6:1 M ratio methanol to oil with 50 mg (0.01 wt.% of
oil) catalyst. The MgO supported K
2
CO
3
catalyst was most efficient
among all the catalyst from different carriers. After 6 cycles, the
catalytic activity decreased minutely but activity was regained
after calcination. The loss of the active sites on the catalyst was
also investigated.
Transesterification of different edible and non-edible oils (such
as sunflower, soybean, ricebran and jatropha) using Mg/Zr catalyst
(catalyst ratio 2:1 wt/wt.%) have been reported by Sree et al., 2009.
Mg/Zr was prepared by co-precipitation method by dissolving
Mg(NO
3
)
2
and ZrO(NO
3
)
2
in deionised water. pH was controlled
at 10 by mixing of two precursors like KOH and K
2
CO
3
. The precip-
itate was filtered then washed and calcined at 650 °C for 4 h. The
XRD results indicated that ZrO
2
was in tetragonal phase, while
MgO was in rocksalt form. The catalyst showed small Zr and large
Mg crystallite sites, making Zr strongly interacted with MgO. How-
ever the high transesterification activity of Mg/Zr catalyst might be
due to the presence of higher number of total basic sites. Total ba-
sicity of catalyst was 1204
l
mol/g while surface area was 47 m
2
/g.
The transesterification reaction was carried out at 65 °C with a
53:1 M ratio of methanol to oil and a catalyst amount of 0.1 g
(0.1 wt.% of oil), to achieve the conversion of about 98% in
50 min. Due to higher number of total basic sites; the high transe-
sterification activity of catalyst was achieved. Insignificant de-
crease of yield up to 5% was observed during transesterification
of sunflower oil after fourth cycle.
Samart et al. (2009) utilized 15 wt.% KI loaded on mesoporous
silica as a solid base catalyst for transesterification of soybean oil
with optimum reaction conditions of 16:1 methanol to oil ratio,
5 wt.% catalyst at 70 °C in 8 h. Conversions of 90% were obtained.
The maximum activity of catalyst was obtained when KI solution
got impregnated on mesoporous silica by incipient wetness
impregnation with concentration of 15 wt.%. The X-ray diffraction
(XRD) patterns of KI/mesoporous silica after calcinations showed
that the characteristic peaks of potassium oxide (K
2
O) face-cen-
tered cubic crystal at 2hequal 25.3°, 41.9°, 51.9°, and 66.9°, while
the characteristic peaks of silicate hydrate phase (SiO
2
xH
2
O) is at
21.8°and 35.7°but there was no characteristic peak of KI in the
XRD pattern because all of KI phases were transformed into K
2
O
phase.
Alumina-supported potassium iodide catalyst was applied for
biodiesel synthesis from soybean oil. The catalyst was prepared
by impregnation of powdered alumina with an aqueous solution
of KI, 35 wt.% KI loaded on Al
2
O
3
and calcined at 500 °C for 3 h
has best catalytic activity and highest basicity (1.5607 mmol/g)
(Xie and Li, 2006). The catalyst activity was dependent on
strength of basic sites as well as upon their amount. On
comparison of alumina loaded with KI, KF, KOH, K
2
CO
3
, KBr and
KNO
3
, the order of conversion reported by the authors was
KI/Al
2
O
3
>KF/Al
2
O
3
>KOH/Al
2
O
3
>KNO
3
/Al
2
O
3
>K
2
CO
3
/Al
2
O
3
>KBr/
Al
2
O
3
.
NaX zeolite loaded with 10% KOH (KOH/NaX) was reported as a
base catalyst in soybean oil transesterification performed by Xie
et al. (2007). NaX zeolite was first dried at 110 °C for 2 h then
impregnated with aqueous solution of KOH for 24 h followed by
drying and by heating at 120 °C for 3 h. The reaction was per-
formed at reflux temperature (65 °C), 10:1 M ratio of methanol to
oil and 3 wt.% catalysts. 85.6% conversion was achieved within
8 h. The results obtained by X-ray diffraction analysis showed
striking similarity in XRD pattern between KOH/NaX samples and
parent zeolites. Further, it was observed by SEM results that NaX
zeolite and KOH/NaX catalysts have nearly spherical shape crystal
with size of 2–4
l
m. The basic strengths of the catalyst as observed
was 15.0 < H_ < 18.4 and it’s very likely that higher percentage of
KOH (>10%) resulted in agglomeration of active sites and hence
lowering the surface areas for active components and resulting
lower the catalytic activity. The regenerated catalyst provides the
conversion rate of 84.3%.
Faria et al. (2008) utilized tetramethylguanidine, which is cova-
lently bonded onto silica gel surface (SiG), as solid catalyst for
transesterification of soybean oil with methanol. SiG catalyst was
prepared by suspending activated silica gel in dry xylene and the
new agent silylant (Silylant was prepared by reacted SiCl with tri-
ethylamine). The SEM image showed that SiG catalyst particles had
spherical morphology with average size of about 1
l
m. It was also
observed that in SiG catalyst covalent immobilization of tetra-
methylguanidine onto the silica gel surface existed, which was
confirmed by FTIR,
29
Si and
13
C NMR spectrums. The pore size
and BET surface area was 8.78 ± 0.73 nm and 216.14 ± 34 m
2
g
1
respectively. The reaction was performed with 1.5 g methanol,
10 g oil and catalyst content of 0.5 g (0.05 wt.% of oil) at 80 °C
and after 3 h, 86% of soybean oil was converted to biodiesel. The
yield decreased continuously from 86% to 62% after catalyst is
recycled for 9 times due to activity drop by loss of the weakly at-
tached tetramethylguanidine to the silica surface.
Georgogianni et al. (2009a) studied conversion of used soybean
frying oil over Mg MCM-41, Mg–Al hydrotalcite and K
+
impreg-
nated zirconia catalysts and found that the Mg–Al hydrotalcite
has the greater activity due to higher basicity. After 25 h under
operation conditions of 60 °C, 5 g oil, 65 ml methanol and 0.5 g of
catalyst (0.1 wt.% of oil), 97% of oil was converted to biodiesel.
Mg/Al hydrotalcite catalyst was synthesized by using mixture of
Mg(NO
3
)
2
6H
2
O, Al(NO
3
)
3
9H
2
O and (NH
4
)
2
CO
3
at 65 °C for 1 h.
The pH was controlled at 5 by concentrate ammonia solution.
The mixture was agitated for 3 h at 65 °C and precipitate was fil-
tered, dried and the calcined at 500 °C for 3 h.
Georgogianni et al. (2009b) also tested Mg MCM-41, Mg–Al
hydrotalcite and K
+
impregnated zirconia catalysts and found that
2154 S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161
Mg–Al hydrotalcite was more active catalyst due to its highly basic
nature for transesterification of rapeseed oil giving quantitative
yield. MCM-41 also gave high yields of methyl esters in the transe-
sterification reaction (conversion 87%). Authors also compared
these results with that of homogenous catalyst (NaOH) in the
transesterification reaction under identical reaction conditions
and found that the homogenous catalyst accelerated the transeste-
rification reaction significantly and gave the equivalent conversion
only within 15 min.
KF/ZnO catalyst has been reported as solid base catalyst for the
transesterification of palm oil with methanol to produce biodiesel
(Hameed et al., 2009). The KF/ZnO was synthesized by impregna-
tion of the ZnO support with aqueous of KF2H
2
O followed by over-
night drying at 110 °C and calcination at 600 °C for 5 h. A loading of
35 wt.% of KF was done on ZnO and good conversions were
achieved.
Yan et al., 2009 studied the ZnO-La
2
O
3
catalyst for transesterifi-
cation of unrefined and waste oil. ZnO-La
2
O
3
was prepared by
homogeneous co-precipitation method where 2 M Zn(NO
3
)
2
and
1 M La(NO
3
) solutions were prepared in distilled water. These solu-
tions, with various Zn–La ratios, were mixed with a 2 M urea solu-
tion and the resulting mixture calcined at 450 °C for 8 h. The
catalyst with 3:1 ratio of Zinc to lanthanum was found to exhibit
highest activity in the transesterification of unrefined or waste
oil. A strong interaction between Zn and La species was observed
with enhanced catalytic activities. The catalyst was active in both
transesterification and esterification reactions, and with no hydro-
lytic activity.
Sr(NO
3
)
2
/ZnO catalyst has been reported by Yang and Xie, 2007
for soybean oil transesterification with methanol at 65 °C.
Sr(NO
3
)
2
/ZnO was prepared by an impregnation method with an
aqueous solution of an alkaline earth metal nitrate and calcined
at 600 °C for 5 h. The optimum catalytic activity was obtained by
loading 2.5 mmol of strontium nitrate Sr(NO
3
)
2
/g on ZnO. The ba-
sicity of catalyst was 10.8 mmol/g and 94.7% soybean oil conver-
sion was achieved after 5 h with 5 wt.% of Sr(NO
3
)
2
/ZnO and 12:1
methanol/oil molar ratio. However, Sr(NO
3
)
2
/ZnO after recovery
exhibited lower catalytic activity with a conversion of soybean
oil of 15.4% and it’s basicity decreased from 10.32 to 6.79 mmol/
g. This was due to decomposition of reactants and products on
the active sites and their interactions during the reaction. But the
catalytic activity of reused catalyst was regained by impregnating
it in an aqueous solution of Sr(NO
3
)
2
. They also observed that the
co-solvent such as dimethyl sulfoxide (DMSO), n-hexane and tetra-
hydrofuran (THF) overcomes the mixing problems of transesterifi-
cation system and among them THF was the most effective co-
solvent which increased the conversion rate of soybean oil up to
96.8%.
The activity and selectivity of NaOH/
c
-Al
2
O
3
catalyst for the
transesterification of sunflower oil with methanol was investigated
by Arzamendi et al. (2007). NaOH/
c
Al
2
O
3
was synthesized by
incipient wetness impregnation of the 212–300
l
m size fraction
of alumina. Prior to use alumina support was calcined at 500 °C
for 12 h. Subsequently, the required amount of NaOH solution
were slowly added to the support, dried for 12 h at 120 °C followed
by calcination at 400 °C for 12 h. NaOH contents of the final solids
were 10.7 and 19.3 wt.% and this was referred as 10-Al and 19-Al
respectively. XRD analysis revealed that a calcined and non-cal-
cined 19-Al catalyst was very similar with the presence of NaOH
and Na
2
O
2
diffraction peaks, both as hydrated compounds, as well
as sodium aluminate (NaAlO
2
). These results indicate that NaOH
has reacted with the support giving rise to the formation of alumi-
nate. For calcined sample, the reaction was performed at 50 °C,
12:1 M ratio of methanol/oil with 19-Al at 0.4 wt.% of NaOH which
gave the conversion rate of about 86% for 24 h. While the conver-
sion for the non-calcined 19-Al sample increased up to 99%. Thus
the result indicated that calcination of NaOH/
c
Al
2
O
3
catalyst had
a negative effect on their activity.
Benjapornkulaphong et al. (2009) compared the catalytic per-
formance of Al
2
O
3
–supported alkali, alkali earth metal oxides and
effect of calcination temperature on activity of different catalyst
for transesterification of palm kernel oil and crude coconut oil with
methanol. They found that Ca(NO
3
)
2
/Al
2
O
3
calcined at 450 °C was
the most suitable catalyst giving 94.3% conversion, however when
the calcination temperature was increased the methyl ester forma-
tion dropped due to the formation of inactive metal aluminates. On
the other hand NaNO
3
/Al
2
O
3
and KNO
3
/Al
2
O
3
improved methyl es-
ter formation tendency at the calcination temperature of above
550 °C but LiNO
3
/Al
2
O
3
catalyst was active with conversion of
91.6% at 450 °C calcination and 93.4% at 550 °C. Mg(NO
3
)
2
/Al
2
O
3
catalyst was not active at any calcination temperature (conversion
10.4% at 450 °C). These catalysts were prepared by the incipient
wetness impregnation of aqueous solution of the corresponding
metal salt precursors (nitrate salt of alkali and alkali earth metals)
on an aluminum oxide support followed by calcination at 450 °C
for 2 h. XRD results indicated that after calcination at 450 °Cof
Ca(NO
3
)
2
/Al
2
O
3
catalyst, mainly CaO species were formed. But
when the calcination was performed at higher temperature, crys-
talline CaO peak decreased. After 3 h of reaction time, at 60 °C with
65:1 M ratio of alcohol/oil and 10 wt.% catalyst content, the maxi-
mum conversion achieved was 94.3% from palm kernel oil whereas
only 85% conversion was obtained in the case of crude coconut oil
due to high acid value and moisture content of crude coconut oil
than palm kernel oil. But when catalyst amount was increased
from 15 to 20 wt.%, the conversion of crude coconut oil also in-
creased from 94% to 99.8%. Results of experiment are tabulated
in Table 1.
Kumar et al. (2010) reported an effective catalyst composition
which contained major amount of nickel zinc aluminate supported
on clay and alumina. On a continuous reactor system, this catalyst
gave conversion in the range of 40–60% at 200 °C and 40 bar
pressure.
Abdullah et al., 2009 reported use of SBA-15 as a neutral mate-
rial made up of Si-O-Si network as a catalyst for biodiesel produc-
tion. The incorporation of potassium into mesoporous SBA-15 (K/
SBA-15) imparted basicity to make it suitable for base catalyzed
reaction like transesterification of palm oil. The authors used a
composite experimental design for optimization of biodiesel yield
so that this mathematical model could predict the biodiesel yield
at any point of time in the experimental domain as well as the
determination of the optimal biodiesel conditions with sufficient
degree of accuracy. K/SBA-15 was prepared by impregnating the
high surface area material (mesoporous SBA-15) with 20 wt.%
KOH solution for 24 h. After that, catalyst was dried followed by
calcination at 350 °C for 3 h. The specific surface area of the cata-
lyst was high and has relatively easy diffusion of reactants in the
mesopores and thereby allowing the use of the mesoporous cata-
lyst at high temperature without suffering major structure modifi-
Table 1
Methyl ester content from palm kernel oil over supported metal oxides calcined at
different temperature (**Reaction conditions: 60 °C, 10 wt.% of catalyst amount,
methanol: oil molar ratio 65:1 and 3 h of reaction time).
Catalyst Optimum calcination
temperature (°C)
Methyl ester
content (wt.%)**
Al
2
O
3
450 0
LiNO
3
/
c
Al
2
O
3
550 93.4
NaNO
3
/
c
Al
2
O3 650 95.1
KNO
3
/
c
Al
2
O
3
550 94.7
Mg(NO
3
)
2
/
c
Al
2
O
3
450 10.4
Ca(NO
3
)
2
/
c
Al
2
O
3
450 94.3
S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161 2155
cation. 93% conversion was achieved in 5 h at 70 °C with methanol/
oil molar ratio of 11.6 and catalyst content of 3.91 wt.%.
2.2. Acidic solid catalysts
Di Serio et al. (2007) investigated application of vanadyl phos-
phate (VOP) as catalyst in the transesterification of soybean oil.
They found that the catalyst was active in transesterification reac-
tion with 80% methyl ester yield obtained only after 1 h reaction
time even though the specific surface area of catalyst was low
(2–4 m
2
/g). Vanadyl phosphate was prepared from the suspension
of V
2
O
5
in diluted phosphoric acid and then calcined at 500 °C for
2 h. The catalytic activity was increased by the increasing the cal-
cination temperature, which helped in removing the hydration
water of the sample and thus, increasing the concentration of the
coordinatively unsaturated VO group and resulted in increased Le-
wis acidity of solids. XRD results indicated that uncalcined VO-
PO
4
2H
2
O has crystallographic
a
form of peaks while another
crystallographic
a
II
peaks was observed in the case of calcined
VOPO
4
(VOP
500
).
Shu et al. (2009) studied carbon based solid acid catalyst for
transesterification of cottonseed oil with methanol. The carbon
based solid acid catalyst was prepared by the sulfonation of car-
bonized vegetable oil asphalt. The catalyst was prepared from car-
bonized vegetable oil asphalt and concentrated H
2
SO
4
solution.
Both these ingredients were heated at 210 °C in an oil bath for
10 h. The suspension was diluted by de-ionized water and dried
at 120 °C for 4 h to obtain the sulfonated vegetable oil asphalt cat-
alyst. The authors also compared transesterification efficiency of
asphalt-based catalyst with sulfonated multi-walled carbon nano-
tubes (s-MWCNTs) and observed that the asphalt-based catalyst
showed higher activity than the s-MWCNTs with 90% conversion.
This is because of high Bronsted acid site density (2.21 mmolg
1
),
loose irregular network and large pores (43.90 nm), which provide
more acid sites for the reactants. The low surface area
(7.48 m
2
g
1
) and high SO
3
H density (2.21 mmol g
1
) of the dry
sulfonated carbon catalysts indicated that most of the SO
3
H
groups were in the interior of the catalyst. The sulfonated polycy-
clic aromatic hydrocarbons provided an electron-withdrawing
function to keep the acid site stable. The morphology of carbon cat-
alyst (SEM) indicated that after the sulfonation treatment by con-
centrated H
2
SO
4
, the particle agglomerates had disintegrated to
some extent and the pores have became larger. The disintegration
of the agglomerated particles also implied that the prepared car-
bon material catalyst had a high quantity of external acid sites
which can be made available to the reactant. The catalyst was able
to catalyze the transesterification reaction for two cycles without
any treatment, but from third cycle, activity decreased due to
swelling effect. In recycling experiments combined influence of
catalyst swelling and deactivation due to leaching of SO
3
H can
be seen and further more it was observed that activity improved
when swelling exceeded the effect of the leaching of SO
3
H group.
Sulfated zirconia solid acid catalyst was studied for transesteri-
fication reaction of soybean oil and simultaneous esterification of
oleic acid with methanol and ethanol in a high pressure reactor
by Garcia et al. (2008). Sulfated zirconia was prepared by either
solvent free method (S-ZrO
2
) or standard precipitation method
(SZ). In solvent free method ZrOCl
2
8H
2
O and (NH
4
)
2
SO
4
are mixed
in molar ratio of 1:6 for 20 min at room temperature and calcined
at 600 °C for 5 h. Whereas, in standard precipitation method, SZ
was prepared by precipitation of zirconium oxychloride hydrate
(ZrOCl
2
.8H
2
O) with ammonium hydroxide at pH 8.5 and then
washed, dried and after that powder was sulfated by impregnated
H
2
SO
4
and then calcined at 650 °C for 4 h. They found that sulfated
zirconia prepared by solvent free method was very active in the
transesterification as well as esterification reaction. The conversion
in alcoholysis catalyzed by S-ZrO
2
obtained under optimized con-
ditions at 120 °C, 5 wt.% of catalyst (w/w) was 98.6% (methanoly-
sis) and 92% (ethanolysis) respectively after 1 h. The performance
of ethanolysis was not as good as in methanolysis due to the higher
water content of ethanol (0.44%) compared to methanol (0.08%). S-
ZrO
2
was an amorphous material while SZ are crystalline (tetrago-
nal and monoclinic phases of zirconia) as depicted by XRD.
Carma et al. (2009) studied the Al-MCM-41 mesoporous
molecular sieves with Si/Al ratio of 8 for esterification of palmitic
acid with methanol, ethanol and isopropanol. The catalyst
Al-MCM-41 with ratio of 8 of Si/Al, produced the highest conver-
sion at 130 °C, 0.6 wt.% catalysts with alcohol to acid molar ratio
of 60 and reaction time of 2 h. The conversion rates for catalyst
were 79%, 67%, and 59% for methanol, ethanol, and isopropanol,
respectively. Catalyst Al-MCM-41 was synthesized by dissolving
aluminum chloride hexahydrate (AlCl
3
.6H
2
O) in a cetyltrimethyl-
ammonium (CTABr) and sodium hydroxide (NaOH) under
intense agitation and by adding tetraethylorthosilicate (TEOS).
The final product was then dried in an oven at 105 °C for 24 h, fol-
lowed by thermal treatment in oven at 550 °C for 7 h and thus,
eliminating the surfactant residue (CTABr) from the pores of the
aluminosilicate. The catalyst was calcined at 480 °C for 3 h. XRD re-
sult showed that the samples have crystallographic patterns char-
acteristic of the mesoporous solid aluminosilicate Al-MCM-41. An
increase in the amount of aluminum incorporated into the mesop-
ores leads to disorder in the structural arrangement of Al- MCM-
41.
Zeolite beta modified with La (La/zeolite beta) had been tested
as a solid acid catalyst for methanolysis of soybean oil (Shu et al.,
2007). Zeolite beta has a high silica zeolite, containing an intersect-
ing three dimensional structure of 12 member ring channels. Due
to this relatively voluminous channel structure, it is possible to
carry out numerous acid catalyzed reactions effectively. The La/
zeolite beta catalyst was prepared by an ion exchange method by
the suspension of zeolite beta in lanthanum nitrate (La(NO
3
)
2
aqueous solution under vigorous stirring at room temperature for
3 h and dried at 100 °C for 24 h and finally calcined at 250 °C for
4 h. The conversion of triglyceride of 48.9 wt.% was observed.
The SEM result indicated that crystal structure of La/zeolite beta
became considerably less agglomerated than zeolite beta due to
the modification because of La
3+
(Na
+
is exchanged with La
3+
so
that instable framework aluminum will be stabilized). The higher
Si/Al ratio can be related to higher structure stability of the tetra-
hedral framework aluminum. FTIR result indicated La/Zeolite beta
shows higher conversion and stability than zeolite beta for the pro-
duction of biodiesel, which may be correlated to the higher quan-
tity of external Bronsted acid sites available for the reactants (the
increment of the intensity of Bronsted acid sites can be attributed
to the presence of Si–OH–La groups and La–OH groups in the La/
zeolite beta after ion exchange).
Karmee and Chadha (2005) used Hb-zeolite, montmorillonite K-
10 and ZnO catalysts for transesterification of non-edible oil of
crude Pongammia Pinnata with 1:10 M ratio of oil/methanol,
0.575 g (0.115 wt.% of oil) catalyst in 5 g oil at 120 °C. They found
that ZnO gave the highest conversion rate of 83%, while Hb-zeolite,
montmorillonite K-10 catalyst gave low conversion rates of 59%
and 47% respectively after 24 h of reaction time.
Jitputti et al. (2006) compared the catalyst activities of several
acidic and basic solids catalysts such as ZrO
2
, ZnO, SO
2
4
=SnO
2
,
SO
2
4
=ZrO
2
, KNO
3
/KL zeolite and KNO
3
/ZrO
2
for transesterification
of crude palm kernel oil (PKO) and crude coconut oil (CCO) with
methanol. Among the catalysts, sulfated zirconia (SO
2
4
=ZrO
2
), a
super acid gave the highest amount of methyl ester yield due to
its high acid strength. The reaction was carried out at 200 °C with
1 wt.% catalyst content, 50 bar pressure and 6:1 of methanol/oil ra-
tio in a high pressure reactor. After 1 h reaction time, 90.3 and 86.3
2156 S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161
wt.% of methyl ester was obtained in the crude palm kernel oil and
crude coconut oil respectively. While in comparison ZrO
2
gave the
lower amounts of methyl esters content and yield. Crude palm ker-
nel oil yielded higher methyl ester than crude coconut oil due the
higher free fatty acid and water content of crude coconut oil which
reduced the methyl ester yield. The activity of solid catalysts
for crude palm kernel oil transesterification was SO
2
4
=ZrO
2
>
SO
2
4
=SnO
2
> ZnO > KNO
3
/ZrO
2
> KNO
3
/KL zeolite > ZrO
2
. In the
case of crude coconut oil the catalysts activity was in order
of SO
2
4
=ZrO
2
>SO
2
4
=SnO
2
> ZnO > KNO
3
/KL zeolite > KNO
3
/ZrO
2
>
ZrO
2
. However as compared with SO
2
4
=SnO
2
and SO
2
4
=ZrO
2
, the
basic ZnO catalyst gave higher methyl ester contents (98.9%) but
a lower methyl ester yield (86.1%) in PKO. In addition, spent
SO
2
4
=ZrO
2
, was not directly reused for transesterification (yield
only 27.7 wt.%), as catalyst deactivated due to combination of cat-
alyst leaching and blocking of active sites by the products or unre-
acted starting materials.
Application of sodium molybdate (Na
2
MoO
4
) was reported by
Nakagaki et al. (2008) for the methanolysis of different types of lip-
ids derived from soybean oil such as refined soybean oil (0.7 mg
KOH/gm acid value), degummed soybean oil (1.0 mg KOH/gm acid
value, 180 ppm of phosphorous as phosphatides) and used frying
oil (1.5 mg KOH/gm acid value). The reaction was carried out at
65 °C with 54:1 of methanol/oil ratio, 5 wt.% catalyst contents in
3 h. The conversion achieved for refined soybean oil, degummed
soybean oil and used frying oil were 95.6 wt.%, 92.6 wt.% and
94.6% respectively. The catalytic activity of the compound was
attributed to the presence of the sites of molybdenum (VI) that
has high Lewis acidity and can polarize at the alcohol O–H bond
leading to a transient species, which has high nucleophilic charac-
ter. Na
2
MoO
4
was synthesized by treating MoO
3
at 550 °C for 2 h
with NaOH solution. Subsequently MeOH was added and Na
2-
MoO
4
H
2
O filtered, washed by methanol and acetone and dried at
120 °C for 3 h.
Furuta et al. (2004) studied solid superacid catalysts such as
sulfated tin oxide (STO), tungstated zirconia-alumina (WZA) and
sulfated zirconium-alumina (SZA) for transesterification of soy-
bean oil with methanol at 300 °C with 4.0 g of catalyst: molar ratio
of methanol to oil was 40:1. For the preparation of tungstated
zirconia-alumina (WZA), a mixture of hydrated zirconia powder
(amorphous), hydrated alumina (pseudo-boehmite), aqueous
ammonium metatungstate solution and de-ionized water were
put into a kneader with stirring for 25 min and there after extruded
in cylindrical pellets shape and followed by drying at 130 °C and
calcination at 800 °C for 1 h. The authors observed that among
the catalysts, tungstated zirconia-alumina catalyst showed highest
activity for transesterification with 94% conversion in 8 h reaction
time while sulfated tin oxide and sulfated zirconium-alumina gave
80% and 70% conversions respectively.
Peng et al. (2008) prepared SO
2
4
/TiO
2
–SiO
2
solid for the produc-
tion of biodiesel from low cost feedstocks (50% oleic acid + 50% re-
fined cotton seed oil) with high FFAs in autoclave reactor at 200 °C,
with molar ratio of methanol to oil 9:1 and 3 wt.% catalyst concen-
tration. The 92% conversion was obtained within 70 min reaction
time. SO
2
4
/TiO
2
–SiO
2
catalyst was synthesized when SiO
2
powder
was slowly added to tetraisopropyl titanate solution of isopropyl
alcohol under reflux for 4 h and dried at 110 °C for 2 h and then cal-
cined at 450 °C for 4 h. The subsequent TiO
2
–SiO
2
particles were
soaked in H
2
SO
4
for 1 day and then dried. SO
2
4
/TiO
2
–SiO
2
was fi-
nally obtained after calcination at 500 °C for 4 h. The authors ob-
served that the large specific surface area of catalyst (258 m
2
/g)
and the average pore diameter (10.8 nm) of the catalyst was big
enough for reactant and product molecules to pass through the
channels. The effect of FFA amount on the yield of esters was stud-
ied by adding 10, 30, 50 and 80 wt.% oleic acid to refined cotton-
seed oil under similar reaction conditions and it was observed
that the FFA content increased the yield of methyl ester and the
rate of esterification of oleic acid was higher than the rate of
transesterification of cottonseed oil due to the better solubility of
FFAs of cottonseed oil in methanol.
Zirconia supported tungsten oxide (WO
3
/ZrO
2
) has been tested
as a solid acid catalyst for esterification of palmitic acid with meth-
anol (Ramu et al., 2004). The catalyst was prepared by impregna-
tion of zirconium hydroxide gel with ammonium meta tungstate
with 2.5–25 wt.% WO
3
loading. The catalyst was dried and finally
calcined at 500 °C. The maximum conversion of 98% was obtained
at 5 wt.% WO
3
/ZrO
2
catalyst in 6 h reaction time. The presence of
crystalline WO
3
and monoclinic phase of zirconia appeared to re-
duce the catalytic activity. The acidity of 5 wt.% WO
3
/ZrO
2
catalyst
was 1.04 mmol/g, which decreased by increasing the amount of
WO
3
due to excess coverage of WO
3
species on ZrO
2
.
Lopez et al., 2007 studied tungstated zirconia (ZrO
2
=WO
2
3
)as
strong solid acid catalyst for both esterification and transesterifica-
tion with methanol as a reactant. The authors evaluated the effect
of calcination temperature (400–900 °C) on the catalytic properties
of tungusted zirconia. Catalytic activities of esterification and
transesterification were increased with the formation of polymeric
W species in the presence of the tetragonal phase of the ZrO
2
sup-
port. They concluded that the optimum calcination temperature
800 °C was efficient to activate tungstated zirconia for both transe-
sterification and esterification reactions. They examined the
transesterification of liquid-phase of triacetin at 60 °C and esterifi-
cation of acetic acid at 60 °C (liquid phase) and 120 °C (gas phase)
with methanol. The maximum catalytic activity was obtained with
catalyst calcined at 800 °C due to the Bronsted acid sites which
contribute most of the activity.
Sreeprasanth et al. (2006) reported Fe–Zn double–metal cya-
nide (DMC) complex as a solid acid catalyst for esterification/
transesterification of sunflower oil. Double-metal cyanide com-
plexes have zeolite- like cage structures (Graverau and Garnier,
1984). The catalyst was synthesized by mixing three solutions;
aqueous solution of K
4
Fe(CN)
6
3H
2
O, a solution of ZnCl
2
in mixture
of distilled water and tert-butanol and a solution of tri-block
copolymer, in mixture of water and tert-butanol. The catalyst
was hydrophobic and contained only Lewis acidic sites. This is be-
cause of coordinatively unsaturated Zn
2+
ions in the structure of
the Fe–Zn complex. The Fe–Zn complexes had a spherical morphol-
ogy as shown by scanning electron microscopy (SEM). The transe-
sterification reaction took place at temperature of 170 °C, with
methanol/oil ratio of 15:1 and 3 wt.% of catalyst and after 8 h of
reaction time and 98.3 wt.% conversions was obtained. The catalyst
was compatible for both esterification (high amount of FFA in the
oil) and transesterification. The water content did not influence
the FAME yield due to hydrophobicity of surface. The catalyst
was reused without any purification and no significant drop of
activity was detected in the transesterification reaction.
2.3. Acid–base solid catalysts
Cheaper feedstocks like waste oils, animal fats cannot be con-
verted to biodiesel using the conventional base mediated process,
as the FFA of oils creates the problems of saponification. Acids
can esterify FFA but the slow rates and limitation of using expen-
sive metallurgy makes it less accepted. Heterogeneous catalysts
having both acidic and basic sites have been investigated which
could esterify FFA and at the same time transesterify triglycerides
to biodiesel.
Lin et al. (2006) reported synthesis of mixed metal oxide meso-
porous silica material for TG transesterification and simultaneous
esterification of FFA. They prepared these mesoporous calcium sil-
icate mixed metal catalysts having different amount of calcium
oxide. A co-condensation method was used for preparation in
S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161 2157
which cetyltrimethyl ammonium bromide (CTAB) provided the mi-
celles template in a NaOH catalyzed reaction of tetraethylorthosil-
icate (TEOS) and the metal oxide. The catalyst after isolation was
freed from surfactant CTAB by calcination at 600 °C for 6 h. SEM/
TEM showed increased structural disorder with increasing content
of calcium oxide. XRD analysis showed total absence of peaks asso-
ciated with CaO and solid state NMR showed the structure similar
to crystalline calcium silicate. These catalysts could esterify soy-
bean oil in methanol in 24 h (80 °C) and could also esterify the free
acids. The recovered catalysts could be reused 30 times for transe-
sterification and 8 times for esterification without significant loss
of catalyst activity.
Lin et al. (2008) obtained a patent for preparation of mesopor-
ous calcium, magnesium silicate and barium silicate by co-conden-
sation method. By forming a mixed oxide from strong basic metal
oxide and weak acidic silica, the acidity of silica was significantly
enhanced. In calcium silicates mixed oxide, silica sites were lewis
acidic, Ca sites as basic and hydroxyl group on surface acted as
Brönsted acids. The co-condensation procedure adopted were sim-
ilar to the one reported by authors earlier (Lin et al., 2006). The
three catalysts, having different Ca/Si ratios, were able to transe-
sterify soybean oil in 90–100% conversion level. The effective tem-
perature range was claimed to be 80 °C and complete conversion
took more than 26 h. Under similar conditions complete esterifica-
tion of poultry fat acids could be achieved in 24 h. All the catalysts
were evaluated for recyclability’s and no loss of activity was no-
ticed in 20 cycles.
Very recently Macario et al. (2010) reported a biodiesel produc-
tion process by homogeneous/ heterogeneous catalyst system of
acid–base type. First the acid catalyst, both strong acid type USY,
BEA and weak acid catalyst of the type MCM-41 were prepared
by hydrothermal synthesis procedures. Later, for preparation of
acid–base type catalyst, potassium (K) was loaded on different
materials by ionic exchange methods. For K loading, the calcined
catalyst materials were treated with 1 M KCl solution at 80 °C
and the ratio of solid/solution was kept at 0.01 g/mL. These K
loaded samples were calcined again at 300 °C for 8 h. Transesteri-
fication reactions were carried out at 100 to 180 °C, molar ratio
of oil to methanol at 1:20, and using 5 wt.% of catalyst. At the
end of reaction, the catalyst was separated by centrifugation,
washed with water and dried overnight at 120 °C. It was observed
that strong acid catalysts like USY, BEA were not good for triglyc-
eride conversion and commercial potassium silicate was found to
be much better. The K loading of MCM-41 increased the conversion
of triglyceride to a great extent but biodiesel production was low
as the main products were FFA (32%), mono-glycerides (42%). The
K loaded delaminated zeolites (K ITQ-6) gave 97% triglyceride con-
version and biodiesel yield of 80%, under the similar reaction con-
ditions. However, when the recovered catalyst was recycled, a
sharp decrease in biodiesel yield was observed and this has been
attributed to leaching of K from the catalyst. The authors proposed
a conceptual flow sheet of a continuous process in which two fixed
bed reactors were employed, one for transesterification and the
other for the catalyst regeneration.
2.4. Enzymatic catalysts
Considering the problems of saponification during the transe-
sterification process, of oil having FFA, by adopting the basic cata-
lyst and slow reaction rate in acid catalyzed reactions, large efforts
have been made to investigate the enzymatically catalyzed transe-
sterification of oils. Enzymatic transesterification avoids soap for-
mation, works at neutral pH, lower reaction temperatures and
thus can be economical. The reusability of enzymes by immobiliz-
ing these on solid supports have provided a new window of oppor-
tunity. Several methods for enzymatic immobilization like
covalent bonding, cross-linking and micro-encapsulation have been
reported. Lipase has been the main enzyme used for transesterifica-
tion, as these are cheaper and are able to catalyze both hydrolysis
and transesterification of triglycerides at very mild conditions and
thus are considered for biodiesel production (Goncalves et al.,
1996; Huge-Jensen et al., 1988; Oliveira et al., 1997).
Catalytic behavior of the Rhizomucor miehei lipase (RML) immo-
bilized on zeolite materials has been studied by Macario et al.
(2007) for biodiesel synthesis with olive oil, containing 76 wt.%
of oleic acid, and methanol. The result shows that biocatalysts have
high capabilities to transesterify fatty acids in olive oil for several
cycles with higher total biodiesel productivity compared to using
free enzyme. The results indicated that the zeolitic materials, hav-
ing a large number of Si-OH groups, are able to adsorb the lipase
enzyme in its open conformation. Silicalite-1 obtained by different
synthesis routes (synthesized in alkaline system and fluorine med-
ia i.e. S1 and F-S2) and delaminated zeolite ITQ-2 has been pre-
pared as lipase-supports. The results were compared with free
enzyme and lipase covalently attached to the functionalized sepi-
olite/AlPO
4
. For the synthesis of enzyme immobilization, the RML
enzyme and the calcined support (wt. ratio: free enzyme/support
equal to 2.5) were mixed in 0.2 M phosphate buffer pH 7, and stir-
red at 250 rpm for 24 h at 0 °C. The support with immobilized li-
pase was separated by filtration, washed with de-ionized water
and dried at 25 °C overnight. The total protein concentration was
calculated by UV absorption at 280 nm. The transesterification
reaction were carried out at 40 °C temperature, 5:1 ratio of meth-
anol to oil (Ma and Hanna, 1999), 0.6 g for lipase/S1 and lipase /
ITQ-2 catalysts, 0.4 g for lipase/sepiolite/AlPO
4
or 100 mg of free li-
pase for 3 h reaction time. It was evaluated that the oleic acid con-
version of Lipase/ S1 and Lipase/ITQ-2 and free lipase were about
100%, 91% and 100%. Whereas, the lower oleic acid conversion
(about 65%) and methyl oleate content (about 43%) was obtained
of Lipase/sepiolite/AlPO
4
, but the stability of the lipase/sepiolite/
AlPO
4
biocatalyst was higher than that of the lipase/S1 and lipase/
TQ-2 biocatalysts. The authors also examined the productivity (mg
of methyl oleate/mg of enzyme/h of reaction) of catalysts and ob-
served that catalysts prepared by adsorption show the highest pro-
ductivity (more than twice) than free enzyme. However, the lipase/
sepiolite–AlPO
4
catalyst had a lower productivity than free lipase
as the covalent binding forces reduced the catalytic activity of
the immobilized lipase. The enzyme immobilized on zeolites was
recycled several times but it gradually leached from the support.
The methyl oleate content of biocatalysts was drastically decreased
from first cycle to third cycle. The methyl oleate content decreased
from 79% to 51% for lipase/S1, and from 70% to 43% for lipase/ ITQ-
2. However, the lipase/sepiolite/AlPO
4
catalyst does not show any
enzyme leaching due to covalent binding of enzyme and support.
The recovered catalyst was washed with n-hexane, dried at room
temperature and stored at 0 °C until subsequent use.
Macario et al. (2009) reported that encapsulation of lipase en-
zyme (Rhizomucor miehei lipase) in highly ordered mesoporous
matrix by a sol–gel method that involves the hydrolysis/ polycon-
densation of a silica precursor at neutral pH and room tempera-
ture. The enzyme is encapsulated within the micellar phase of
the surfactant that is self-assembled with silica. The encapsulated
biocatalyst has been used for the transesterification reaction of tri-
olein with methanol under solvent free conditions. The highest
fatty acid methyl esters yield (77%) was obtained after 96 h at
40 °C, with triolein:methanol molar ratio of 1:3 and 5 wt.% of cat-
alyst (1.5 wt.% of enzyme). For the preparation of heterogeneous li-
pase enzyme, two different immobilization procedures such as
encapsulation and the adsorption procedure were studied. In
encapsulation procedure, lipase solution was added to the cetyl-
trimethylammonium bromide (CTMABr) solution and stirred for
1 h at room temperature. The silica precursor was then introduced
2158 S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161
into the solution and, subsequently, ethanolamine (20 wt.%) was
added. The gelation was slow and the sol–gel was stirred for 24 h
at room temperature and pH 7.2 followed by the filtration. After
that the liquid was analyzed by UV-adsorption at 280 nm to deter-
mine the degree of enzyme encapsulation and the enzyme/silica
weight ratio in the final solid catalyst (immobilization yield). In
adsorption-immobilization procedure, 50 ml of a 0.2 M phosphate
buffer solution (pH 7.0) and RML was added to the support in pow-
der form. The mixture was stirred (250 rpm) for 24 h at room tem-
perature. The support with adsorbed lipase was washed twice with
de-ionized water and dried at 25 °C overnight and then the total
protein concentration was measured by UV absorption method at
280 nm. Various enzyme contents of MCM-41 were prepared by
this method and the authors observed that the immobilization
yield of lipase turns out to be higher than 95% for all the samples.
The order of the mesoporous structure moderately decreases when
the enzyme content increases, with molar ratio of enzyme to silica
ranging from 0.005 to 0.020. The pore diameter and BET surface
area and pore diameter were increased by increasing the enzyme
amount that produces the swelling of the micelle of surfactant.
The yield of methyl esters and the enzyme activity increases with
the amount of lipase loaded. It was observed that the FAME yields
of free lipase at 18 h of reaction are lower than those obtained with
the encapsulated lipase. But after 70 h, the FAME yield is close to
80% for the encapsulated and 50% for free lipase. Therefore, due
to interaction of hydrophobic chains of the surfactant with the
hydrophobic patches of the enzyme, enzyme structure gets opened
and accessibility of the lipase catalytic centre to the reactants gets
increased. Hence, even after the presence of surfactant, enzyme
activities are not inhibited. Total productivity of the immobilized
enzyme is almost six times higher than the one obtained using free
lipase.
Caballero et al. (2009) studied the free and immoblized Pig pan-
creatic lipase (PPL) enzyme on sepiolite for transesterification of
sunflower oil and alcohol. The optimum reaction conditions of free
and immobilized Pig pancreatic lipase were: reaction temperature
of 40 °C, oil to ethanol ratio of 2:1 v/v, pH of 12 and catalyst con-
tents of 0.01 g (0.1 wt.% of total substrate) for free PPL and 0.5 g
of demineralised sepiolite containing 0.01 g of immobilised PPL
(0.1 wt.% of total substrate) for immobilized PPL. The PPL activity
was increased on increasing pH value (12) and the maximum bio-
diesel yield found after 10 h reaction time was around 57.7% and
26.9% for free PPL and immobilized PPL respectively. The enzyme
PPL was immobilized on sepiolite, which is a natural silicate having
fibrous structure, after acid treatment to remove Mg atoms. The
Mg free sepiolite was treated with PPL enzyme in ethanol at 0 °C
for 24 h, centrifuged to remove the non-immobilized enzyme.
Through the immobilized enzyme was less efficient as compared
to the free enzyme, but the ease of recyclability and retention of
initial enzyme activity were major factors favour of its use for bio-
diesel production.
Immobilized Lipase (Thermomyces lanuginosus) on novel micro-
porous polymer matrix (MPPM) has been tested for the transesteri-
fication reaction of sunflower, soybean and waste cooking oils with
methanol as a low cost biocatalyst (Dizge et al., 2009). Poly HIPE
using styrene, divinylbenzene, and polyglutaraldehyde was used
to synthesize Microporous polymeric matrix (MPPM) containing
aldehyde functional group. Thermomyces lanuginosus lipase was
covalently attached onto MPPM with 80%, 85%, and 89% immobili-
zation efficiencies using bead, powder, and monolithic forms,
respectively. MPPM synthesis (monolithic, bead, and powder
forms), microporous polymeric biocatalyst (MPPB) preparation by
immobilization of lipase onto MPPM and biodiesel production by
MPPB are the three aspect of the process on which research is fo-
cused. MPPM was prepared by polymerizing the continuous phase
of a high internal phase emulsion consisting of organic and water
phases. The organic phase was composed of styrene, divinylben-
zene, and Span 80. While potassium persulphate and polyglutaral-
dehyde solution were contained in water phase. Immobilization of
lipase was carried out by reaction of powder (4–8 mesh in size) or
bead matrix with enzyme in calcium acetate buffer (25 mM, pH 6)
at 26 °C for 25 h with gentle shaking (250 rpm). Then, to remove the
unbound enzyme the immobilized enzyme was washed with ace-
tate buffer. SEM micrographs and FTIR spectrum showed that
copolymer can be produced as a porous structure having aldehyde
functional groups. The immobilization efficiencies obtained using
bead and powder forms were 80% and 85%, respectively. The transe-
sterification reaction was carried out in accordance with design of
experiment based on Taguchi methodology at 65 °C, 1:6 M ratio
of oil and methanol, 250 rpm of stirring and 0.0108 wt.% of immo-
bilized lipase (powder or bead form) for 24 and 5 h in batch reactor.
Methanol was added to the mixture in three-steps to avoid strong
methanol inhibition. In 5 h reaction time, biodiesel yields for sun-
flower oil was 63.8%, 81.1%, and 86.9% using monolithic, bead,
and powdered MPPB, respectively. It was observed that the most
effective biocatalyst was powdered MPPB for the production of bio-
diesel as it get efficiently mixed with reactants during reaction. It
was also observed that the immobilized enzyme retained the activ-
ity during 10 repeated batch reactions.
3. Conclusions
Several solid acidic catalysts have been investigated for biodie-
sel synthesis but their uses have been limited due to lower reaction
rates and unfavorable side reactions. Basic heterogeneous catalysts
have also been investigated but their activity gets degraded in the
presence of water. Acid–base catalysts are one of the potential cat-
alysts because they catalyze both esterification and transesterifica-
tion simultaneously. Enzymatic catalysts though highly promising
but are rather slow. For a successful commercial catalyst, catalyst
life, recyclability and lower cost are extremely important as these
have a direct effect on overall cost of the process. Only few reports
indicate the commercial level production of biodiesel by adopting
the heterogeneous catalyst route.
Acknowledgements
The authors would like to express their gratitude to the Univer-
sity of Petroleum and Energy Studies, Dehradun and to Manage-
ment of R&D Centre, Indian Oil Corporation Limited, Faridabad
for the permission to publish this work.
One of the author (S. Semwal) would like to thank University of
Petroleum and Energy Studies for award of research fellowship.
References
Abdullah, A.Z., Razali, N., Lee, K.T., 2009. Optimization of mesoporous K/SBA-15
catalyzed transesterification of palm oil using response surface methodology.
Fuel Process. Technol. 90, 958–964.
Abreu, F.R., Lima, D.G., Hamu, E.H., Einloft, S., Rubim, J.C., Suarez, P.A.Z., 2003. New
metal catalysts for soybean oil transesterification. J. Am. Oil Chem. Soc. 80, 601–
604.
Arzamendi, G., Campo, I., Arguinarena, E., Sanchez, M., Montes, M., Gandia, L.M.,
2007. Synthesis of biodiesel with heterogeneous NaOH/alumina catalysts:
comparison with homogeneous NaOH. Chem. Eng. 134, 123–130.
Arzamendi, G., Arguinarena, E., Campo, I., Zabala, S., Gandia, L.M., 2008. Alkaline and
alkaline-earth metals compounds as catalysts for the methonolysis of sunflower
oil. Catal. Today, 133–135.
Azam, M.M., Waris, A., Nahar, N.M., 2005. Prospects and potential of fatty acid
methyl esters of some non-traditional seed oils for use as biodiesel in India.
Biomass Bioenergy 29, 293–302.
Benjapornkulaphong, S., Ngamcharussrivichai, C., Bunyakiat, K., 2009. Al
2
O
3
-
supported alkali earth metal oxides for transesterification of palm kernel oil
and coconut oil. Chem. Eng. J. 145, 468–474.
Caballero, V., Bautista, F.M., Campelo, J.M., Luna, D., Marinas, J.M., Romero, A.A.,
Hidalgo, J.M., Luque, R., Macario, A., Giordano, G., 2009. Sustainable preparation
S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161 2159
of a novel glycerol-free biofuel by using Pig pancreatic lipase: partial 1, 3-
regiospecific alcoholysis of sunflower oil. Process Biochem. 44, 334–342.
Canakci, M., 2007. The potential of restaurant waste lipids as biodiesel feedstocks.
Bioresour. Technol. 98, 183–190.
Cantrell, D.G., Gillie, L.J., Lee, A.F., Wilson, K., 2005. Structure reactivity correlations
in MgAl hydrotalcite catalysts for biodiesel synthesis. Appl. Catal. A: Gen. 287,
183–190.
Carma Jr., A.C., de Souza, L.K.C., da Costa, C.E.F., Longo, E., Zamian, J.R., da Rocha
Filho, G.N., 2009. Production of biodiesel by esterification of palmitic acid over
mesoporous aluminosilicate Al-MCM-41. Fuel 88, 461–468.
Dalai, A.K., Kulkarni, M.G., Meher, L.C., 2006. Biodiesel productions from vegetable
oils using heterogeneous catalysts and their applications as lubricity additives.
IEEEEIC Climate Change Technology Conference, EICCCC art 4057358.
Di Serio, M., Cozzolino, M., Tesser, R., Patrono, P., Pinzari, F., Bonelli, B., Santacesaria,
E., 2007. Vanadyl phosphate catalyst in biodiesel production. Appl. Catal. A:
Gen. 320, 1–7.
Di Serio, M., Tesser, M.R., Pengmei, L., Santacesaria, E., 2008. Heterogeneous
catalysts for biodiesel production. Energy Fuels 22, 207–217.
Dizge, N., Aydiner, C., Imer, D.Y., Bayramoglu, M., Tanriseven, A., Keskinler, B., 2009.
Biodiesel production from sunflower, soybean, and waste cooking oils by
transesterification using lipase immobilized onto a novel microporous polymer.
Bioresour. Technol. 100, 1983–1991.
Dorado, M.P., Ballesteros, E., Anal, J.M., Gomez, J., Lopez, F.J., 2003. Exhaust
emissions from a diesel engine fueled with transesterified waste olive oil. Fuel
82, 1311–1315.
Dossin, T.F., Reyniers, M.F., Berger, R.J., Marin, G.B., 2006. Simulation of
heterogeneously MgO-catalyzed transesterification for fine-chemical and
biodiesel industrial production. Appl. Catal. B: Gen. 67, 136–148.
Faria, E.A., Ramalho, H.F., Marques, J.S., Suarez, P.A.Z., Prado, A.G.S., 2008.
Tetramethylguanidine covalently bonded onto silica gel surface as an efficient
and reusable catalyst for transesterification of vegetable oil. Appl. Catal. A: Gen.
338, 72–78.
Furuta, S., Matsuhashi, H., Arata, K., 2004. Biodiesel fuel production with solid
superacid catalysis in fixed bed reactor under atmospheric pressure. Catal.
Commun. 5, 721–723.
Garcia, C.M., Teixeira, S., Marciniuk, L.L., Schuchardt, U., 2008. Transesterification of
soybean oil catalyzed by sulfated zirconia. Bioresour. Technol. 99, 6608–6613.
Georgogianni, K.G., Katsoulidis, A.K., Pomonis, P.J., Manos, G., Kontominas, M.G.,
2009a. Transesterification of rapeseed oil for the production of biodiesel using
homogeneous and heterogeneous catalysis. Fuel Process. Technol. 90, 1016–
1022.
Georgogianni, K.G., Katsoulidis, A.P., Pomonis, P.J., Kontominas, M.G., 2009b.
Transesterification of soybean frying oil of biodiesel using heterogeneous
catalysts. Fuel Process. Technol. 90, 671–676.
Goncalves, A.P.V., Lopes, L.M., lemos, F., Ramoa Ribeiro, F., Prazeres, D.M.F., Cabral,
J.M.S., Aires-Barros, M.R., 1996. Zeolites as support for enzymatic hydrolysis
reactions. Comparative study of several zeolites. J. Mol. Catal. B: Enzym. 1, 53–
60.
Granados, M.L., Poves, M.D.Z., Alonso, D.M., Mariscal, R., Galisteo, F.C., Tost, R.M.,
et al., 2007. Biodiesel from sunflower oil by using activated calcium oxide, Appl.
Catal. B: Environ. 73, 317–326.
Graverau, P., Garnier, E., 1984. Structure de la phase cubique de l’hexacyanoferrate
(III) de zinc : Zn
3
[Fe(CN)
6
]
2
. Acta Crystallogr. C 40, 1306–1309.
Hameed, B.H., Lai, L.F., Chin, L.H., 2009. Production of biodiesel from palm oil (Elaeis
guineensis) using heterogeneous catalyst: an optimized process. Fuel Process.
Technol. 90, 606–610.
Huge-Jensen, B., Galluzzo, D.R., Jensen, R.G., 1988. Studies on free and immobilized
lipases from Mucor miehei. J. Am. Oil Chem. Soc. 65, 905–910.
Ji, J., Wang, J., Li, Y., Yu, Y., Xu, Z., 2006. Preparation of biodiesel with the help of
ultrasonic and hydrodynamic cavitation. Ultrasonics 44, 411–414.
Jitputti, J., Kitiyanan, B., Rangsunvigit, P., Bunyakiat, K., Attanatho, L.,
Jenvanitpanjakul, P., 2006. Transesterification of crude palm kernel oil and
crude coconut oil by different solid catalysts. Chem. Eng. J. 116, 61–66.
Karmee, S.K., Chadha, A., 2005. Preparation of biodiesel from crude oil of Pongamia
pinnata. Bioresour. Technol. 96, 1425–1429.
Kawashima, A., Matsubara, K., Honda, K., 2008. Development of heterogeneous base
catalysts for biodiesel production. Bioresour. Technol. 99, 3439–3443.
Kawashima, A., Matsubara, K., Honda, K., 2009. Acceleration of catalytic activity of
calcium oxide for biodiesel production. Bioresour. Technol. 100, 696–700.
Kouzu, M., Kasuno, T., Tajika, M., Yamanaka, S., Hidaka, J., 2008. Active phase of
calcium oxide used as solid base catalyst for transesterification of soybean oil
with refluxing methanol. Appl. Catal. A: Gen. 334, 357–365.
Kumar, S., Pulikottil, A.C., Sharma, A., Sarkar, M., Kagdiyal, V., Yadav, A., Arora, A.K.,
Puri, S.K., Patel, M.B., Santra, M., Sarin, R., Kumar, B., Gupta, A.A., Basu, B.,
Malhotra, R.K., Kumar, A., 2010. A catalyst composition for transesterification of
organically/ naturally derived oils and fats to produce alkyl esters and process
for preparing the same. WO 2010/020998 A2.
Kwiecien, J., Hajek, M., Skopal, F., 2009. The effect of the acidity of rapeseed oil on its
transesterification. Bioresour. Technol. 100, 5555–5559.
Li, H., Xie, W., 2006. Transesterification of soybean oil to biodiesel with Zn/I
2
catalyst. Catal. Lett. 107, 25–30.
Liang, X., Gao, S., Wu, H., Yang, J., 2009. Highly efficient procedure for the synthesis
of biodiesel from soybean oil. Fuel Process. Technol. 90, 701–704.
Lin, V.S.-Y., Nieweg, J.A., Kern, C., Trewyn, G., Wiench, J.W., Pruski, M., 2006. Acid-
base mesoporous calcia-silica catalysts for cooperative conversion of bio-based
feedstocks into biodiesel. Prepr. Symp. Am. Chem. Soc. 51, 426–427.
Lin, V.S-Y., Nieweg, J.A., Verkade, J.G., Venkat, R., Reddy, C., Kern, C., 2008. New
composite based catalysts for biodiesel production. Patent WO 2008/013551
A1.
Liu, X., He, H., Wang, Y., Zhu, S., 2007. Transesterification of soybean oil to biodiesel
using SrO as a solid base catalyst. Catal. Commun. 8, 1107–1111.
Liu, X., He, H., Wang, Y., Zhu, S., Ziao, X., 2008. Transesterification of soybean oil to
biodiesel using CaO as a solid base catalyst. Fuel 87, 216–221.
Lopez, D.E., Suwannakarn, K., Bruce, D.A., Goodwin Jr., J.G., 2007. Esterification and
transesterification on tungstated zirconia: effect of calcination temperature. J.
Catal. 247, 43–50.
Ma, F., Hanna, M.A., 1999. Biodiesel production: a review. Bioresour. Technol. 70, 1–
15.
Macario, A., Giordano, G., Setti, L., Parise, A., Campelo, J.M., Marinas, J.M., Luna, D.,
2007. Study of lipase immobilization on zeolitic support and transesterification
reaction in a solvent free-system. Biocatal. Biotransform. 25, 328–335.
Macario, A., Moliner, M., Corma, A., Giordano, G., 2009. Increasing stability and
productivity of lipase enzyme by encapsulation in a porous organic- inorganic
system. Microporous Mesoporous Mater. 118, 334–340.
Macario, A., Giordano, G., Onida, B., Cocina, D., TAGARELLI, A., Giuffre, A.M., 2010.
Biodiesel production process by homogeneous/ heterogeneous catalystic
system using an acid-base catalyst. Appl. Catal.A; Gen. 378, 160–168.
Mbaraka, I.K., Shanks, B.H., 2006. Conversion of oils and fats using advanced
mesoporous heterogeneous catalysts. J. Am. Oil Chem. Soc. 83, 79–91.
Meher, L.C., Sager, D.V., Naik, S.N., 2006. Technical aspects of biodiesel production
by transesterification- a review. Renew. Sustain. Energy Rev. 10, 248–268.
Minami, E., Saka, S., 2006. Kinetics of hydrolysis and methyl esterification for
biodiesel production in two-step supercritical methanol process. Fuel 85, 2479–
2483.
Nakagaki, S., Bail, A., dosSantos, V.C., deSouza, V.H. R., Vrubel, H., Nunes, F.S., Ramos,
L.P., 2008. Use of anhydrous sodium molybdate as an efficient heterogeneous
catalyst for soybean oil methanolysis. Appl. Catal. A: Gen. 351, 267–274.
Oliveira, A.C., Rosa, M.F., Cabral, J.M.S., Aires-Barros, M.R., 1997. Immobilization of
saccharomyces cerevisiae cells and Rhizomucor miehei lipase for the production
and extractive biocatalysis of ethanol. Bioprocess Eng. 16, 349–353.
Peng, B.X., Shu, Q., Wang, J.F., Wang, G.R., Wang, D.Z., Han, M.H., 2008. Biodiesel
production from waste oil feedstock by solid acid catalysis. Process Saf. Environ.
Prot. 86, 441–447.
Ramu, S., lingaiah, N., Devi, B.L.A.P., Prasad, R.B.N., Suryanarayana, I., Prasad, P.S.S.,
2004. Esterification 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. 276,
163–168.
Samart, C., Sreetongkittikul, P., Sookman, C., 2009. Heterogeneous catalysis of
transesterification of soybean oil using KI/mesoporous silica. Fuel Process.
Technol. 90, 922–925.
Sarin, R., Sharma, M., Sinharay, S., Malhotra, R.K., 2007. Jatropha-palm biodiesel
blends: an optimum mix for Asia. Fuel 86, 1365–1371.
Sarin, R., Sharma, M., Khan, A.A., 2009. Studies of Guizotia abyssinica L. Oil: biodiesel
synthesis and process optimization. Bioresour. Technol. 100, 4187–4192.
Sarin, R., Arora, A.K., Puri, S.K., Prakash, S., Ranjan, R., Christopher, J., Tuli, D.K.,
Malhotra, R..K., Kumar, A., 2009b. Novel catalyst composition for biodiesel
production and a process for producing biodiesel and product thereof. Patent
applied (Appl.No. 1348IN033).
Sharma, M., Kumar, R., Ray, S.S., Malhotra, R.K., Sarin, R.,Verma, R.P., Reje, N.R., 2006.
Process for producing biodiesel and product thereof. US Patent 2006/0094890
A1.
Shimada, Y., Watanabe, Y., Sugihara, A., Tominaga, Y., 2002. Enzymatic alcoholysis
for biodiesel fuel production and application of the reaction to oil processing. J.
Mol. Catal. B: Enzym. 17, 133–142.
Shu, Q., Yang, B., Yuan, H., Qing, S., Zhu, G., 2007. Synthesis of biodiesel from
soybean oil and methanol catalyzed by zeolite beta modified with La
3+
. Catal.
Commun. 8, 2159–2165.
Shu, Q., Zhang, Q., Xu, G., Nawaz, Z., Wang, D., Wang, J., 2009. Synthesis of biodiesel
from cottonseed oil and methanol using a carbon-based solid acid catalyst. Fuel
Process. Technol. 90, 1002–1008.
Sree, R., Babu, N.S., Prasad, P.S.S., Lingaiah, N., 2009. Transesterification of edible and
non-edible oils over basic solid Mg/Zr catalysts. Fuel Process. Technol. 90, 152–
157.
Sreeprasanth, P.S., Srivastava, R., Srinivas, D., Ratnasamy, P., 2006. Hydrophobic,
solid acid catalysts for production of biofuels and lubricants. Appl. Catal. A: Gen.
314, 148–159.
Suppes, G.J., Dasari, M.A., Doskocil, E.J., Mankidy, P.J., Goff, M.J., 2004.
Transesterification of soybean oil with zeolite and metal catalysts. Appl. Catal.
A: Gen. 257, 213–223.
Tiwari, A.K., Kumar, A., Raheman, H., 2007. Biodiesel production from jatropha oil
(jatropha curcas) with high free fatty acids: an optimized process. Biomass
Bioenergy 31, 569–575.
Veljkovic, V.B., Stamenkovic, O.S., Todorovic, Z.B., Lazic, M.L., Skala, D.U., 2009.
Kinetics of sunflower oil methanolysis catalyzed by calcium oxide. Fuel 88,
554–1562.
2160 S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161
Wang, L., Yang, J., 2007. Transesterification of soybean oil with nano-MgO or not in
supercritical and subcritical methanol. Fuel 86, 328–333.
Xie, W., Huang, X., 2006. Synthesis of biodiesel from soybean oil using
heterogeneous KF/ZnO catalyst. Catal. Lett. 107, 53–59.
Xie, W., Li, H., 2006. Alumina-supported potassium iodide as a heterogeneous
catalyst for biodiesel production from soybean oil. J. Mol. Catal. A: Chem. 255,
1–9.
Xie, W., Huang, X., Li, H., 2007. Soybean oil methyl esters preparation using NaX
zeolites loaded with KOH as a heterogeneous catalyst. Bioresour. Technol. 98,
936–939.
Yan, S., Salley, S.O., Simon Ng, K.Y., 2009. Simultaneous transesterification and
esterification of unrefined or waste oils over ZnO-La
2
O
3
catalysts. Appl. Catal. A:
Gen. 353, 203–212.
Yang, Z., Xie, W., 2007. Soybean oil transesterification over zinc oxide modified with
alkali earth metals. Fuel Process. Technol. 88, 631–638.
Zabeti, M., Daud, W.M.A.W., Aroua, M.K., 2009. Activity of solid catalysts for
biodiesel production: a review. Fuel Process. Technol. 90, 770–777.
S. Semwal et al. / Bioresource Technology 102 (2011) 2151–2161 2161
... [18]. However, a longer reaction time makes this unattractive compared to the short reaction time from the base-catalysed transesterification process [37,38]. Several types of solid base catalysts have been utilized for biodiesel production such as alkaline metal oxide, alkaline metal carbonates or hydro-carbonates, anionic resins and basic zeolites [39]. ...
Article
Full-text available
The dwindling of global petroleum deposits and worsening environmental issues have triggered researchers to find an alternative energy such as biodiesel. Biodiesel can be produced via transesterification of vegetable oil or animal fat with alcohol in the presence of a catalyst. A heterogeneous catalyst at an economical price has been studied widely for biodiesel production. It was noted that various types of natural waste shell are a potential calcium resource for generation of bio-based CaO, with comparable chemical characteristics, that greatly enhance the transesterification activity. However, CaO catalyzed transesterification is limited in its stability and studies have shown deterioration of catalytic reactivity when the catalyst is reused for several cycles. For this reason, different approaches are reviewed in the present study, which focuses on modification of waste-shell derived CaO based catalyst with the aim of better transesterification reactivity and high reusability of the catalyst for biodiesel production. The catalyst stability and leaching profile of the modified waste shell derived CaO is discussed. In addition, a critical discussion of the structure, composition of the waste shell, mechanism of CaO catalyzed reaction, recent progress in biodiesel reactor systems and challenges in the industrial sector are also included in this review.
... Heterogeneous base catalysts derived from solid waste shells, alkaline and alkali earth metals are a better replacement over homogeneous catalysts in biodiesel production by a transesterification process (Syazwani et al. 2017;Niju et al. 2014a, b, c;Girish et al. 2013;Chouhan and Sarma 2011;Trisupakitti et al. 2018). Use of modified solid base heterogeneous catalysts in biodiesel production by the transesterification process has been attempted using various feedstocks in recent times and some of the selected catalysts are cited in the literature (Surbhi et al. 2011;Wang et al. 2013;Wenlei and Haitao 2006;Niju et al. 2014a;Anjana et al. 2016;Sneha et al. 2015). ...
Article
Experimental parameters influencing the transesterification of rubber seed oil (RSO) to biodiesel using alumina (Al 2 O 3) impregnated on calcined eggshells (Al 2 O 3 /calcined eggshells) were studied in the present work. Parameters were optimized using response surface methodology (RSM) and artificial neural network (ANN). A conversion of 98.9% was observed for RSO at optimum conditions of 12:1 methanol: oil molar ratio, 3 (wt%) catalyst concentration and 4 (h) of reaction time. A significant quadratic model with molar ratio as the most influencing process parameter and a coefficient of determination, R 2 , of value equal to 0.9379 is observed from RSM analysis. Best validation performance of 5.8595 at epoch-1 and R 2 value equal to 0.9740 was observed from ANN modeling. On comparing RSM and ANN models, it is concluded that ANN is a better tool for predicting the conversion of RSO to biodiesel with minimum error.
... Heterogeneous base catalysts derived from solid waste shells, alkaline and alkali earth metals are a better replacement over homogeneous catalysts in biodiesel production by a transesterification process (Syazwani et al. 2017;Niju et al. 2014a, b, c;Girish et al. 2013;Chouhan and Sarma 2011;Trisupakitti et al. 2018). Use of modified solid base heterogeneous catalysts in biodiesel production by the transesterification process has been attempted using various feedstocks in recent times and some of the selected catalysts are cited in the literature (Surbhi et al. 2011;Wang et al. 2013;Wenlei and Haitao 2006;Niju et al. 2014a;Anjana et al. 2016;Sneha et al. 2015). ...
Article
Experimental parameters influencing the transesterification of rubber seed oil (RSO) to biodiesel using alumina (Al2O3) impregnated on calcined eggshells (Al2O3/calcined eggshells) were studied in the present work. Parameters were optimized using response surface methodology (RSM) and artificial neural network (ANN). A conversion of 98.9% was observed for RSO at optimum conditions of 12:1 methanol: oil molar ratio, 3 (wt%) catalyst concentration and 4 (h) of reaction time. A significant quadratic model with molar ratio as the most influencing process parameter and a coefficient of determination, R2, of value equal to 0.9379 is observed from RSM analysis. Best validation performance of 5.8595 at epoch-1 and R2 value equal to 0.9740 was observed from ANN modeling. On comparing RSM and ANN models, it is concluded that ANN is a better tool for predicting the conversion of RSO to biodiesel with minimum error.
... It contains no sulfur compounds and hence the chances of SOx production is literally nil or negligible during its combustion (Jai-In et al. 2014). Chemically, biodiesel can be produced by transesterification of triglycerides with alcohol in the presence of a catalyst (Semwal et al. 2011). ...
Article
Application of a novel heterogeneous catalyst fluorite (calcined fluorspar CaF 2) in biodiesel preparation from a highly viscous non-edible rubber seed oil has been studied. Optimization of process parameters and effect of process parameters on biodiesel conversion is done using response surface methodology (RSM) with central composite design analysis. The optimum process conditions, i.e., 12:1 molar ratio (mol/mol), 4 (wt%) catalyst, and 5 h of reaction time for a conversion of 95.95% oil to biodiesel conversion were observed. A quadratic regression model to predict the conversion in terms of process parameters was obtained. Artificial neural network (ANN) with the Levenberg-Marquardt algorithm is also attempted, with the best performance of 0.80239 at epoch-3 and a mean square error of 1.24. On comparison of the coefficient of determination, R 2 , the value obtained from both RSM and ANN it is observed that the ANN model predicts conversion better than RSM, with minimum error. ARTICLE HISTORY
Conference Paper
Full-text available
Dhaka is the capital city of Bangladesh. A lot of people migrate every day. For their survival, the housing sector is flourishing day by day. The demand impacts the supplies. A number of people are doing housing business to support the situation. Still, on the contrary, they are grabbing a lot of green fields or water bodies from private properties evacuating the owners. One of the ill examples is considered as Pallabi Eastern Housing 2nd Phase Project. The study is to find the socio-economic status of the land owners and local inhabitants who are evicted from their land for the project. According to the primary observation and information collected from different sources, approximately 350 families are relocated. Of the 45 relocated people placed under survey from the questionnaire survey by snowball sampling, it was found that most people were displaced from their original location through the involuntary procedure. The income level, housing condition, and utility facilities are worse off than before the resettlement. They are unaware of taking action against the developer for the deprivation in the land price. The study discovered different aspects that clarified how land developers exploited the local people. The four different actors, namely land developers, local musclemen, local officials, and local people, were identified who actively played various roles in this relocation procedure. A few scopes are identified to rehabilitate the victims to recover from their present situation. Government should adequately care about the reallocated people to compensate at the current rate considering the long period of suffering. There is scope for more intensive study on the topic.
Chapter
This study focused on recycling used cooking oil from residential dwellings and making it a promising alternative to fossil fuel and the collected oil was converted into biodiesel and tested on a diesel engine rig. The fuel consumption of biodiesel was observed to be slightly higher compared to standard diesel due to the lower heating value of the biodiesel in addition to its higher viscosity and density. Lower emissions of CO and HC have observed in biodiesel compared to diesel whereas the CO2 emissions were almost the same for both fuels. The NOX emissions were observed to be higher as compared to diesel due to several factors like higher oxygen content in biodiesel, short ignition delay, and advances in injection and combustion of the engines.
Article
Full-text available
In the recent times, biodiesel production from various non‐edible oil sources has received a lot of attention. Karanja oil is chosen mainly due to its availability and low cost. It is a nitrogen fixing tree cultivated to enhance soil quality. Transesterification of Karanja oil to biodiesel using a basic solid modified heterogeneous catalyst has been discussed in the present study. Aluminum nitrate Al(NO3)3 impregnated on calcined marble (Al(NO3)3/calcined marble), prepared by wet impregnation method (25%), is used as a modified heterogeneous base catalyst. Surface morphology of the synthesized catalyst is determined using a scanning electron microscope (SEM) analyzer and the composition of various elements present in the prepared catalyst is analyzed using energy dispersive spectrometer (EDS analyzer). 67.56 (wt/wt) of calcium oxide (CaO) is observed for calcined marble and 78.19 (wt/wt %) of CaO has been noticed for impregnated catalyst (Al(NO3)3/calcined marble). A maximum biodiesel yield of 96.5% and a conversion of 97.43% was obtained at an optimum methanol:oil molar ratio of 9:1 (mol/mol), 4 (wt %) of catalyst concentration and 1.5 (h) reaction time respectively. The entire composition of formed biodiesel is determined using gas chromatography–mass spectrometry (GC–MS) analyzer and conversion of Karanja oil to biodiesel is measured using ¹H‐nuclear magnetic resonance (NMR) analyzer.
Article
The replacement of traditional diesel fuels with an alternative energy source like biodiesel is an excellent choice to lessen the depletion of non-renewable energy source such as fossil fuels. For application of low-cost heterogeneous catalyst for generation of biodiesel, most of waste-based catalysts are derived from inexpensive CaO resources. In this research, heterogeneous catalyst aluminum chloride hexahydrate (AlCl36H2O)/PVA was synthesized from aluminum foil wastes supported by PVA. Transesterification was used to produce biodiesel from waste cooking oil (WCO). The effect of aluminum chloride hexahydrate (AlCl36H2O)/PVA catalyst towards the yield percentage of FAME was investigated using three different operating conditions. Parameters optimized in transesterification included methanol to oil ratio (6:1, 12:1 and 18:1), reaction temperature (55℃, 60℃, and 65℃), and amount of catalyst (2 wt%, 5 wt%, and 8 wt%). The findings showed that the catalyst loading of 2 wt%, methanol to oil molar ratio of 6:1, and temperature of 55℃ yielded the highest biodiesel yield from WCO. This study proves that heterogeneous catalyst aluminum chloride hexahydrate (AlCl36H2O)/PVA can replace homogeneous catalysts and to simplify the manufacturing of biodiesel, especially in the separation phase.
Article
The main objective of the present work is to work on optimization of process parameters in acid value minimization of biodiesel produced from rubber seed oil (RSO) using Al2O3/Calcined Eggshells as a heterogeneous catalyst. Optimization of process parameters was performed using response surface methodology (RSM) and artificial neural network (ANN). Acid value of raw RSO is reduced from 67.6 (mg KOH/g oil) to 2.97 (mg KOH/g oil) and later it has been reduced to 0.33 (mg KOH/g oil) in the synthesized biodiesel. Form the ANOVA analysis of RSM studies, it is observed that the designed quadratic model and the parameter methanol: oil molar ratio (mol/mol) are significant. ANN model with best validation performance of 0.00065251 at epoch-2 is observed. On comparison of coefficient of regression R2 value of 0.8885 for RSM studies and 0.9939 for ANN studies it was noticed that ANN is the best fit model with minimum error.
Article
Full-text available
Silver fish (Mene maculata) is a type of pelagic fish found in the Moluccas sea, this fish has a high oil content so that it is easy to extract the oil content for transesterification. The transesterification reaction of oil from silver fish oil needs a catalyst, and heterogenous catalyst is a better choice. This study was determined the fatty acid methyl ester composition of M. maculata by transesterification process using heterogeneous CaO catalyst from chicken egg shells through 3 steps i.e. oil isolation, CaO catalyst synthesis, and oil transesterification reactions. Oil isolation was carried out by soxhletation method using petroleum ether with the oil content was 13.51%. Synthesis of CaO catalyst was carried out by calcination at 1000 °C for 2 hours and then charactererized by XRD. Fish oil is transesterified using CaO catalyst in a ratio of 11% (w/v) of the mixture. Furthermore, fatty acid methyl esters obtained were analyzed using GC-MS. The results showed 6 methyl ester fatty acids as composition of transesterification products which were methyl myristic (11.85%), methyl palmitoleate (8.71%), methyl palmitic (25.13%), methyl oleic (9.49%), methyl stearic (10.04%), and methyl 5,8,11,14-eikosatetraenoate (1.96%).
Article
Full-text available
The methanolysis of sunflower oil was studied in the presence of CaO previously calcined at various tem- peratures and the optimal temperature for CaO calcination was determined. The sigmoidal process kinet- ics was explained by the initial triglyceride (TG) mass transfer controlled region, followed by the chemical reaction controlled region in the latter reaction period. The TG mass transfer limitation was due to the small available active specific catalyst surface, which was mainly covered by adsorbed mole- cules of methanol. In the later phase, the adsorbed methanol concentration decreased, causing the increase of both the available active specific catalyst surface and the TG mass transfer rate, and the chem- ical reaction rate become smaller than the TG mass transfer rate.
Article
Full-text available
Zn/I2 was found to be a practical and effective catalyst for the transesterification of soybean oil with methanol. A study for optimizing the reaction conditions such as the molar ratio of methanol to oil, the reaction time and the catalyst amount, was performed. The highest conversion of 96% was obtained under the optimum conditions. Further, the effect of free fatty acids and water in the soybean oil on the catalytic activity of the catalysts was also investigated.
Article
The production of biodiesel is greatly increasing due to its enviromental benefits. However, production costs are still rather high, compared to petroleum-based diesel fuel. The introduction of a solid heterogeneous catalyst in biodiesel production could reduce its price, becoming competitive with diesel also from a financial point of view. Therefore, great research efforts have been underway recently to find the right catalysts. This paper will be concerned with reviewing acid and basic heterogeneous catalyst performances for biodiesel production, examining both scientific and patent literature.
Article
Sodium molybdate (Na2MoO4) has been synthesized and investigated as a heterogeneous catalyst for the methanolysis of different types of renewable lipid sources derived from the soybean oil. Transesterification reactions occurred under relatively mild conditions, requiring low temperatures, short times and normal pressure. The methyl esters synthesized were characterized by reverse-phase and size-exclusion chromatography and hydrogen-nuclear magnetic resonance spectroscopy. The catalyst was characterized by X-ray diffraction, thermogravimetric analysis, and Fourier transform infrared spectroscopy (FTIR). It is very efficient for the transesterification reaction of triglycerides with methanol, with yields higher than 95% of methyl esters. The molybdenum(VI)-complex has a high Lewis acidity and most certainly act on alcohol O–H bond leading to a transient species which has high nucleophilic character. The catalyst was easily recovered and after being washed showed capacity of recyclability for another catalytic reaction with similar activity.
Article
In this work, we report the encapsulation of lipase enzyme in highly ordered mesoporous matrix by a sol–gel method that involves the hydrolysis/polycondensation of a silica precursor at neutral pH and room temperature. The enzyme is encapsulated within the micellar phase of the surfactant that is self-assembled with silica. The encapsulated biocatalyst has been used for the transesterification reaction of triolein with methanol under solvent free conditions. The highest fatty acid methyl esters yield (77%) was obtained after 96h at 40°C, with triolein:methanol molar ratio equal to 1:3 and with a very low amount of enzyme. Total productivity of the immobilized enzyme is almost six times higher than the one obtained using free lipase. The results clearly indicate that the immobilization procedure of lipase preserves the mobility of the enzyme and allows to increase its stability.
Article
As part of ongoing efforts to investigate heterogeneous catalysts for biodiesel synthesis, the catalytic activity of the strong solid acid tungstated zirconia (WZ) was studied for the esterification of acetic acid (in both the gas and liquid phase) and the transesterification of triacetin (a synthetic triglyceride in the liquid phase) with methanol. Acetic acid and triacetin were used as model compounds for free fatty acids (FFAs) and triglycerides, respectively. Complex mixtures of FFAs and triglycerides are typical biodiesel precursors found in unrefined vegetable oils, animal fats, and waste greases. The effect of calcination temperature (400–900 °C) on the catalytic properties of WZ was investigated with the goal of determining the optimum pretreatment temperature and increasing the understanding of the nature of the active sites for esterification and transesterification. Catalyst characterization by powder X-ray diffraction, Raman spectroscopy, TGA, elemental analysis, and total BET surface area allowed us to correlate the changes in catalyst physicochemical properties with calcination temperature. Ion-exchange/titration and NH3-TPD were used to characterize the number and strength of the acid sites. Esterification and transesterification rates (on a catalyst weight basis) increased with increasing surface WOx densities between 1.9 and 6.6 W-atoms nm−2 and then decreased for densities above 6.6 W-atoms nm−2. The growth of WO3 crystalline species resulted in lower esterification and transesterification rates, consistent with a decrease in the number of active sites. For the catalyst that exhibited the optimum catalytic activity (calcined at 800 °C), selective poisoning of the potential catalytic centers revealed that Brønsted sites play the major role in carrying out these reactions.
Article
La/zeolite beta was prepared by an ion exchange method and used to synthesize the biodiesel (fatty acid methyl esters, FAME). The La(NO3)3 was applied as the ion exchange precursor to incorporate La ion into zeolite beta. The composition of the zeolite beta before and after ion exchange was analyzed by the SEM microphotographs and EDS spectrograms, the Brønsted and Lewis acid sites were investigated by FTIR imaging. The transesterification was carried out in a batch reactor and the composition of the FAME product was determined by a potassium hydroxide saponification method. The syntheses conditions with respect to catalytic activities have been optimized individually. Results of the experiment showed that La/zeolite beta shows higher conversion and stability than zeolite beta for the production of biodiesel, which may be correlated to the higher quantity of external Brønsted acid sites available for the reactants. The product consists of a mixture of monoalkyl esters primarily, and when the methanol/ soybean oil molar ratio was 14.5, reaction temperature at 333K, reaction time 4h and catalyst/soybean oil mass ratio of 0.011, the conversion of triglyceride 48.9wt% was obtained from this optimal reaction condition.