Anda di halaman 1dari 13

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/233578345

Catalysts in Production of Biodiesel: A Review


ARTICLE in JOURNAL OF BIOBASED MATERIALS AND BIOENERGY MARCH 2007
Impact Factor: 0.65

CITATIONS

READS

65

781

3 AUTHORS:
Narasimharao Katabathini

Adam F Lee

King Abdulaziz University

Aston University

56 PUBLICATIONS 582 CITATIONS

211 PUBLICATIONS 4,431 CITATIONS

SEE PROFILE

SEE PROFILE

Karen Wilson
Aston University
228 PUBLICATIONS 4,748 CITATIONS
SEE PROFILE

Available from: Narasimharao Katabathini


Retrieved on: 17 November 2015

Copyright 2007 American Scientic Publishers


All rights reserved
Printed in the United States of America

Journal of
Biobased Materials and Bioenergy
Vol. 1, 112, 2007

REVIEW

Catalysts in Production of Biodiesel: A Review


K. Narasimharao , Adam Lee, and Karen Wilson
Department of Chemistry, University of York, York YO10 5DD, UK
Biodiesel is a renewable substitute fuel for petroleum diesel fuel which is made from nontoxic,
biodegradable, renewable sources such as rened and used vegetable oils and animal fats.
Biodiesel is produced by transesterication in which oil or fat is reacted with a monohydric alcohol
in the presence of a catalyst. The process of transesterication is affected by the mode of reaction,
molar ratio of alcohol to oil, type of alcohol, nature and amount of catalysts, reaction time, and temperature. Various studies have been carried out using different oils as the raw material and different
alcohols (methanol, ethanol, butanol), as well as different catalysts, notably homogeneous ones
such as sodium hydroxide, potassium hydroxide, sulfuric acid, and supercritical uids or enzymes
such as lipases. Recent research has focused on the application of heterogeneous catalysts to
produce biodiesel, because of their environmental and economic advantages. This paper reviews
the literature regarding both catalytic and noncatalytic production of biodiesel. Advantages and disadvantages of different methods and catalysts used are discussed. We also discuss the importance
of developing a single catalyst for both esterication and transesterication reactions.

Keywords:

CONTENTS
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1. Biodiesel Usage Facts . . . . . . . . . . . . . . . . . .
1.2. Biodiesel Feed Stocks . . . . . . . . . . . . . . . . . .
2. Biodiesel Production . . . . . . . . . . . . . . . . . . . . . .
2.1. Base Catalyzed Transesterication . . . . . . . . . .
2.2. Acid-Catalyzed Transesterication . . . . . . . . . .
2.3. Enzyme Catalysts . . . . . . . . . . . . . . . . . . . . .
2.4. Supercritical Transesterication . . . . . . . . . . . .
2.5. Esterication of Free Fatty Acids . . . . . . . . . . .
2.6. Simultaneous Transesterication and Esterication
2.7. Biodiesel from Vegetable Oil via Pyrolysis . . . . .
3. Future Scope of Research . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

1
2
3
4
4
6
7
7
8
9
9
10
10
11

1. INTRODUCTION
The depletion of world petroleum reserves and increased
environmental concerns have stimulated the search for
alternative renewable fuels that are capable of fullling
an increasing energy demand.1 In recent decades, research
concerning and knowledge about the external benets of
renewable raw materials have intensied the efforts for
sustainable energy sources. Biodiesel plays a major role
in this eld because of the world-wide research, development, and deployment activities of this sustainable energy
source.2

Authors to whom correspondence should be addressed.

J. Biobased Materials and Bioenergy 2007, Vol. 1, No. 1

Biodiesel fuel (fatty acid methyl ester (FAME)) from


vegetable oil, which primarily contains triglycerides (TGs)
and free fatty acids (FFAs), is considered the best candidate for diesel fuel substitute in diesel engines and is used
neat (100% biodiesel) or can be blended with petroleum
diesel.3 It is an attractive alternative (or extender) to
petroleum diesel fuel due to well-known advantages,
including the following:
(i) it provides the potential for lower dependence on
petroleum crude oil,
(ii) it is a renewable resource,
(iii) it provides the potential for reduced greenhouse gas
emissions because of the closed CO2 cycle (Scheme 1),
(iv) it has a lower combustion emission prole (especially SOx ),
(v) it provides the potential for enhancement of rural
economies,
(vi) it is biodegradable,
(vii) it can be used without engine modications,
(viii) it provides good engine performance,
(ix) improved combustion is exhibited because of its
oxygen content,
(x) it exhibits low toxicity, and nally
(xi) it has the ability to be blended in any proportion with
regular petroleum-based diesel fuel.4
It can also be used in boilers or furnaces designed to use
heating oils or in oil-fueled lighting equipment.
1556-6560/2007/1/001/012

doi:10.1166/jbmb.2007.002

Catalysts in Production of Biodiesel: A Review

Waste Oil

Fatty acid
esters

fica

teri

ses

n
Tra

tion

Narasimharao et al.

Reduced
smog
& CO2

REVIEW

Triglycerides

Rapeseed
or soybean
oil
Carbon
neutral
CO2

Scheme 1. CO2 life cycle for biodiesel.

Due to the recent increased awareness and development


in this area, the objective of this review is to give fundamental insight into the production of biodiesel by different catalytic materials. In an effort to identify the ideal
catalyst characteristics for biodiesel synthesis, this review
also strives to compare the efciency and applicability

of a number of liquid and solid catalysts in the production


of biodiesel.
1.1. Biodiesel Usage Facts
As early as the beginning of the 20th century Rudolf
Diesel proposed vegetable oil as a fuel for his engine.5
During World War Two, vegetable oil was examined in
up-to-date diesel engines, while in 1940, vegetable oil
methyl and ethyl esters were used in France and Belgium
as a fuel for buses. High consumption of conventional
energy resources and increasing emission regulations has
motivated an intense search for alternative fuels over the
decades. Many proposals have been made regarding
the availability and applicability of an environmentally
friendly fuel that could be domestically available.6
Methanol, ethanol, compressed natural gas (CNG), liqueed natural gas (LNG), liqueed petroleum gas (LPG), and
vegetable oils have all been considered as alternative fuels.
In the 1930s and 1940s vegetable oils were used as diesel
fuels from time to time, but usually only in emergency
situations.7 Researchers have also concluded that vegetable

Dr. Narasimharao completed his Master of Science in 1996 from Andhra University, Vizag,
India and a Ph.D. in 2002 from the Indian Institute of Chemical Technology, Hyderabad,
India. After post doctoral positions in the University of Missouri-Kansas City, USA and The
Technical University of Kaiserslautern, Germany he was awarded a Royal Society Indian
visiting fellowship to work at the University of York (200506).

Dr. Lee has a B.A. (Natural Sciences) and Ph.D. from the University of Cambridge and is
currently a Senior Lecturer in Physical Chemistry at the University of York. Prior to this he
was a Lecturer in Physical Chemistry at the University of Hull, and conducted post-doctoral
research at Cambridge and Johnson Matthey Technology Centre. Adam was awarded the
2000 C R Burch prize of the British Vacuum Council and 2004 Fonda-Fasella Elettra prize
for outstanding contributions to vacuum and synchrotron science, respectively.

Dr. Wilson has a B.A. (Natural Sciences) and Ph.D. from the University of Cambridge
and an M.Sc. in catalysis from Liverpool University. Following postdoctoral research both
at Cambridge and within the Green Chemistry Group at the University of York she was
appointed in 1999 to a Lecturership in Physical Chemistry in the Department of Chemistry
at University of York.

J. Biobased Materials and Bioenergy 1, 112, 2007

Narasimharao et al.

Catalysts in Production of Biodiesel: A Review

vehicles and has also been reported to deteriorate polymers, such as polyurethane foam materials.
1.2. Biodiesel Feed Stocks
Biodiesel is dened as the alkyl monoesters of fatty acids
from renewable resources, such as vegetable oils, animal
fats, and waste restaurant oils and greases. Biodiesel development can be found in 28 countries, Germany and France
being the largest producers of biodiesel fuel in the world.12
There are several choices for vegetable oil sources. In
the United States and Brazil soybean oil is a source that is
already scaled up for biodiesel production. Nevertheless,
other sources, such as rapeseed (in Europe), sunower,
peanut, cotton, palm oil, coconut, babassu, and especially
castor oil, may also used in different parts of the world
once their cultivation can achieve an economic up-scaling.4
The alcohol source is generally methanol, though ethanol
and butanol have also been used. In general the physical and chemical properties and the performance of ethyl
esters are comparable to those of the methyl esters. Methyl
and ethyl esters have almost the same heat content, but the
viscosities of the ethyl esters are slightly higher.13 Engine
tests demonstrate that methyl esters produce slightly higher
power and torque than ethyl esters,14 the latter also tending to cause more injector coking than methyl esters. The
other alcohols are less reactive than methanol, and there
are some technological problems in its industrial use.
Different oils and fats contain a range of fatty acid concentrations (Table I), and as a result, biodiesel production
costs are highly dependent on the feedstock type. The cost
of the fat or oil used to produce biodiesel clearly affects
the cost of the nished product, constituting up to 60
75% of the overall nancial burden; therefore, less expensive raw materials are preferred.15 To produce biodiesel
economically, a production facility must have access to
low-value feedstock, develop quality, high-value coproducts, and enjoy a cost-effective and high-yielding process.
Vegetable oils, such as soybean or rapeseed oil, are significantly more expensive than beef tallow or yellow grease,
despite the higher free fatty acid content of the latter.
In the UK recycled cooking oil is currently the main
feedstock of choice for biodiesel production due to its
low purchase price. Once the limited quantities of used

Table I. Fatty acid distribution of some vegetable oils and animal fat.
Fatty acid distribution (% by weight)
Feed stock
Rapeseed oil
Sunower oil
Safower oil
Soybean oil
Beef tallow
Yellow grease

C14:0
36
243

C16:0
349
608
860
1058
2432
2324

C16:1
3.79

J. Biobased Materials and Bioenergy 1, 112, 2007

C18:0

C18:1

C18:2

C18:3

Saturation level (%)

085
326
193
476
2025
1296

64.40
16.93
11.58
22.52
37.43
44.36

2230
7373
7789
5234
23
697

8.23

434
934
1053
1534
4763
3863

8.19
0.67

REVIEW

oils hold promise as replacement fuels for modern diesel


engines since their caloric value is comparable to that of
diesel. However, their use in direct injection diesel engines
is restricted by some unfavorable physical properties, particularly their viscosity, which is about 1117 times higher
than that of diesel fuel. Consequently, vegetable oil causes
poor fuel atomization, incomplete combustion, and carbon
deposition on the injector and valve seats, resulting in serious engine fouling.4
Different approaches have been considered to reduce the
high viscosity of vegetable oils: (a) dilution of 25 parts
vegetable oil with 75 parts diesel fuel, (b) microemulsions
with short chain alcohols, (c) pyrolysis, (d) catalytic cracking, and (e) transesterication with ethanol or methanol
to produce fatty acid esters commonly known as biodiesel
fuel.8
Biodiesel is renewable, nontoxic, and biodegradable.
However, while biodiesel is denitely renewable, the fact
that it cannot displace a signicant fraction of current
petroleum-based fuel consumption means that it does not
really allow us to make much progress toward using it
as a sustainable energy supply. Indeed if all of the available vegetable oil and animal fat were used in biodiesel
synthesis we could only replace about 15% of the current demand for on-highway diesel fuel.9 Nontoxicity and
biodegradability are useful characteristics, but they are
only signicant when the fuel is used in its pure form,
as is common in Germany and Austria. For the 20% and
lower blends that are common in the United States, the
diesel fuel portion of the blend determines the toxicity and
biodegradability.10 Biodiesel does provide a reduction in
harmful emissions, such as SOx , CO, hydrocarbons, particle matter, and soot, as well as NOx in optimized diesel
engines, together with lower net CO2 emissions.
The solvent properties of biodiesel can also cause some
problems when used initially in diesel engines. Sediments
and sludge formed in old diesel storage tanks can become
dispersed upon initial use of biodiesel, resulting in blocked
lters, but once the contaminants are removed no further
maintenance issues arise. Indeed the superior lubricity of
biodiesel can even help reduce engine wear.11 However,
pure biodiesel is not compatible with natural rubber, so it
can degrade some fuel system components made of natural
rubber such as hoses, gaskets, and seals found in pre-1993

Catalysts in Production of Biodiesel: A Review

REVIEW

industrial frying oils have been fully given over to


biodiesel, additional raw feedstocks may come from the
cultivation of rapeseed crops.16

2. BIODIESEL PRODUCTION
Vegetable oils and animal fats are comprised of a complex mixture of triglycerides and other minor components,
such as free fatty acids, gums, waxes, etc. Triglycerides
are esters of glycerol with three chains of aliphatic or
olenic FFAs of variable length (1224 carbons). Direct
use of oil diluted with solvents to form microemulsions
lowers viscosity and improves ignition characteristics but
can result in some engine performance problems.17 The
pyrolysis of vegetable oil has been investigated for over
100 years as a route to synthesize petroleum. While this
tends to produce more biogasoline than biodiesel fuel,18
it is viewed by some companies as an economic method
to produce valuable alternative biofuels. Among all the
proposed methods to convert oils to biodiesel, transesterication of the triglycerides seems to be the best choice, as
the physical characteristics of fatty acid esters (biodiesel)
are very close to those of diesel fuel.19 Furthermore, the
methyl or ethyl esters of fatty acids can be burned directly
in unmodied diesel engines, with very low deposit formation and a by-product (glycerol) that has commercial
value.20
In the biodiesel production process, transesterication
is the chemical reaction between triglycerides and alcohol in the presence of a catalyst to produce monoesters.
The long and branched chain triglyceride molecules are
transformed to monoesters and glycerol. The transesterication process consists of a sequence of three consecutive reversible reactions, which include conversion of
triglycerides to diglycerides, followed by the conversion
of diglycerides to monoglycerides. The glycerides are converted into glycerol and yield one ester molecule in each
step. The overall transesterication reaction can be represented by the reaction Scheme 2.
Several aspects, including the type of catalyst (alkaline
or acid), alcohol/vegetable oil molar ratio, temperature,
purity of the reactants (mainly water content), and free
fatty acid content, have an inuence on the course of the
transesterication.21 The alcohol/vegetable oil molar ratio
is one of the main factors that inuences the transesterication. An excess of the alcohol favors the formation of the
products. On the other hand, an excessive amount of alcohol makes the recovery of the glycerol difcult. The high

Scheme 2. Transesterication of TGs with alcohols.

Narasimharao et al.

molar ratio of alcohol to vegetable oil interferes with the


separation of glycerine, because there is an increase in solubility. When glycerin remains in solution, it helps drive
the equilibrium back to the left, lowering the yield of
esters. So, the ideal alcohol/oil ratio has to be established
empirically, considering each individual process.
There are different transesterication processes that
can be applied to synthesize biodiesel: (a) base-catalyzed
transesterication, (b) acid-catalyzed transesterication,
(c) enzyme-catalyzed transesterication, and (d) supercritical alcohol transesterication.
2.1. Base Catalyzed Transesterication
2.1.1. Homogeneous Catalysts
Biodiesel is currently synthesized using homogeneous
alkaline catalysts because the transesterication reaction
by an acid catalyst is much slower than the base-catalyzed
reaction. There are several comprehensive studies of basecatalyzed transesterication.2123 The most common basic
catalysts are potassium hydroxide (KOH), sodium hydroxide (NaOH), sodium methoxide (NaOCH3 ), and sodium
ethoxide (NaOCH2 CH3 ). Daranoko et al.24 studied the
two-step transesterication reaction of waste oils by using
stoichiometric amounts of methanol and the necessary
amounts of KOH, supplemented with the exact amount of
KOH to neutralize acidity. Both reactions were completed
in 30 min in the temperature range 4060  C. Parameters such as density, viscosity, water content, and energy
content were investigated. It was concluded that a twostep, alkaline-catalyzed transesterication reaction is an
economic method for biodiesel production from used vegetable oil.
The activity and efciency of nonionic bases as catalysts for the transesterication of vegetable oils in homogeneous media has also been studied.25 In a rst study,
the catalytic activity of some guanidines was compared.
It was observed that 1,5,7-triazabicyclo[4.4.0]dec-5-ene
(TBD) produces more than 90% of methyl esters after 1 h
with only 1 mol% in the reaction mixture. The other bases
tested under the same experimental conditions gave yields
below 66%. The catalytic activity of these materials is not
directly related to the relative basicity of compounds. The
catalytic activity of TBD was also compared with that of
typical industrial catalysts and the yields obtained with
TBD were close to those observed with NaOH with no
by-products such as soaps, which are easily formed when
alkaline metal hydroxides are used. When compared to
potassium carbonate, TBD was always more active, even
at low molar concentrations. Although TBD is less active
than sodium methoxide at only 0.5 mol%, it does not
require any special reaction conditions.
Due to the excellent performance of TBD in the transesterication of vegetable oils, catalytic activity of several
alkylguanidines was investigated under the same reaction
conditions (27.2 mmol of rapeseed oil, 62.5 mmol
J. Biobased Materials and Bioenergy 1, 112, 2007

Narasimharao et al.

2.1.2. Heterogeneous Catalysts


Even though homogeneous catalyzed biodiesel production
processes are relatively fast and show high conversions
with minimal side reactions, they are still not very costcompetitive with petrodiesel. For instance, the basic catalysts used in the process, generally KOH and NaOH, are
neutralized with phosphoric acid after the reaction, and the
resultant salts constitute vast quantities of undesired waste
chemicals. The other disadvantages include that
(1) the catalyst cannot be recovered and must be neutralized at the end of the reaction,
(2) there is limited use of continuous processing
methodologies,30 and
(3) the processes are very sensitive to the presence of
water and free fatty acids; consequently they need a high
quality feedstock (rened vegetable oils) to avoid undesired side reactions (hydrolysis and saponication) or additional reaction steps to rst convert/eliminate the free fatty
acids.
Biodiesel synthesis using solid catalysts instead of homogeneous catalysts could potentially lead to cheaper
production costs by enabling reuse of the catalyst and
opportunities to operate in a xed bed continuous
process.31
J. Biobased Materials and Bioenergy 1, 112, 2007

Solid basic materials such as MgO, AlMg hydrotalcites, Cs-exchanged sepiolite, and mesoporous MCM-41
have been used as catalysts for glycerol transesterication
with triglycerides.32 MgO and low Al content hydrotalcites behave as active catalysts for triolein glycerolysis.
Other esters have been used for glycerol transesterication. Metal oxides like MgO, CeO2 , La2 O3 , and ZnO have
been used as solid base catalysts for the transesterication
of glycerol with stoichiometric amounts of methyl stearate
in the absence of solvent.33 These catalysts are active, but
the selectivity to mono-, di-, and triesters is similar to
that obtained by using homogeneous basic catalysts (40%
monoester at 80% conversion).
Schuchardt et al.34 heterogenized guanidines on
organic polymers such as cellulose and poly(styrene/
divinylbenzene) for the transesterication of vegetable
oils. The guanidine-containing cellulose shows a slightly
reduced activity compared to guanidine in the homogeneous phase, giving a conversion of 30% after 1 h,
when used at 5 mol%. This guanidine-containing cellulose was used in a continuous reactor containing 100 g
of the catalyst with a 2:1 alcohol/oil mixture pumped at
60  C at a rate of 0.48 liters/h. However, they observed
an incomplete reaction attributed either to leaching of
the catalyst or to its irreversible protonation. In order to
circumvent the leaching of the guanidines from the polymers, Sercheli et al.35 reported a method to encapsulate
N ,N  ,N  -tricyclohexylguanidine (TCG) in the supercages
of a hydrophobic Y zeolite by the reaction of dicyclohexylcarbodiimide and cyclohexylamine. These encapsulated
guanidines showed low activity in the transesterication of
vegetable oils (14% conversion after 5 h), as diffusion of
the triglycerides through the channels of the Y zeolite is
slowed due to steric hindrance. The guanidines heterogenized on gel-type poly(styrene/divinylbenzene) showed a
slightly lower activity than their homogeneous analogues
but allowed the same high conversions after prolonged
reaction times. However, they also slowly leached from
the polymers, allowing only nine catalytic cycles.36
Insoluble salts of amino acids have also found usage
as catalysts for the methanolysis of triglycerides.37 Some
metal salts of amino acids such as those of copper, zinc,
cadmium, nickel, lanthanum, cobalt, calcium, magnesium, and iron were tested. Zinc arginate catalysts are
also applied for the methanolysis of palm oil with a
methanol:oil molar ratio of 6:1 and deliver 67 wt% of
methyl ester. However, reasonable reaction rates could
only be achieved at temperatures above 130  C, and it is
unclear whether these amino acid catalysts are reusable
or not.
The transesterication of soybean oil using calcium
carbonate as a catalyst has been investigated.38 Conversions above 95% are achieved at 260  C for ethyl esters,
using ow reactors with residence times of approximately
18 min. No decrease in the activity of calcium carbonate
5

REVIEW

of methanol, and 1.0 mol% of catalyst at reaction temperature of 70  C) for the sake of comparison.26 The TBD was
always the most active (90% yield in 1 h reaction time);
however, 1,3-dicyclohexyl-2-n-octylguanidine (DCOG),
1,1,2,3,3-pentamethylguanidine (PMG), 7-methyl-1,5,7triazabicyclo[4.4.0]dec-5-ene (MTBD), and 1,2,3-tricyclohexylguanidine (TCG) also showed yields of 64%, 47%,
49%, and 74% in 1 h reaction time. The activity order of
the catalysts is TBD > TCG > DCOG > MTBD > PMG,
in line with their relative base strengths.
Metal complexes27 28 of the type M(3-hydroxy-2methyl-4-pyrone)2 (H2 O)2 , where M = Sn, Zn, Pb, and Hg,
have been used for soybean oil methanolysis under homogeneous conditions. Sn and Zn complexes showed great
activities for this reaction, achieving yields of up to
90% and 40%, respectively, in 3 h, using a molar ratio
of 400:100:1 (methanol:oil:catalyst), without emulsion
formation.
Recently,29 liquid amine-based catalysts were also
successfully applied for transesterication of rened vegetable and frying oil. Four amines (diethylamine
(DEA), dimethylethanol amine (DMAE), tetramethyldiaminoethane (TEMED), and tertramethylammonium
hydroxide (TMAH) (as 25% in methanol)) were used. The
highest conversion of 98% was achieved with TMAH as a
catalyst at 65  C in 90 min. In these cases, a large amount
(13%) of liquid amine catalysts is required for the transesterication. Amine catalysts used in the reaction also
act as a solvent for both reactants and products, helping
to shift the reaction equilibrium toward the products.

Catalysts in Production of Biodiesel: A Review

Narasimharao et al.

was observed after weeks of utilization, but these catalytic


systems need high energy requirements. The methanolysis
of rapeseed oil was tested in the presence of caesiumexchanged NaX faujasites, mixed magnesiumaluminum
oxides, magnesium oxide, and barium hydroxide as catalysts for different methanol:oil ratios.39 Barium hydroxide was particularly catalytically active for a methanol:oil
molar ratio of 6:1 under methanol reux after a reaction time of only 1 h; oil conversion was about 80%
with a nearly quantitative ester molar fraction. On the
other hand, caesium-exchanged NaX faujasites and mixed
magnesiumaluminum oxides required a long reaction
time (3 h) and a high methanol:oil molar ratio to achieve
high yields in methyl esters. The transesterication of soybean oil with methanol in the presence of a series NaX
faujasite zeolite, titanosilicalites with Na and K as cations
(ETS-10), zeolite, and metal catalysts has also been
studied.40 ETS-10 catalysts offered higher conversions
(>80%) than zeolite-X type catalysts; however, this may
reect a homogeneous route as alkali methoxide species
were leached out. Na/NaOH/-Al2 O3 heterogeneous alkaline catalysts were applied for the transesterication of
soybean oil with methanol using hexane as cosolvent.41
The best result (>90%) was obtained after 2 h using a
methanol:oil molar ratio of 9:1 at 60  C.
We have reported a series of Li-promoted CaO
catalysts42 with Li loadings in the range 0.264.0 wt%
which are effective in the transesterication of glyceryl
tributyrate and methanol to methyl butanoate. A Li content of 1.23 wt% was found to give the optimum activity toward methyl butanoate formation (Fig. 1). Li doping
increases the base strength of CaO, and XPS and DRIFTS
measurements reveal that the optimum loading correlates
with the formation of an electron decient surface Li+
species and associated OH species at defect sites on the
support. High Li loadings result in bulk LiNO3 formation
3
Li (Li/CaO)

Initial rate (mmolmin1g(cat)1)

REVIEW

Catalysts in Production of Biodiesel: A Review

Mg (hydrotalcite)

2.2. Acid-Catalyzed Transesterication


The transesterication reaction can also be catalyzed by
Brnsted acids, preferably sulfonic and sulfuric acids, but
the reactions rates are low and require relatively high temperatures to get high product yields.45 According to an
acid-catalyzed mechanism for esterication,46 carboxylic
acids can be readily formed by hydrolysis of the carbocation intermediate formed upon protonation of the ester.
This suggests that acid-catalyzed transesterication should
be carried out in the absence of water to avoid the competitive formation of carboxylic acids and concomitant reduction in the yields of alkyl esters.

2.2.1. Homogeneous Catalysts

0
0

10

12

Loading (mmolg1)
Fig. 1. Catalyst activity as a function of Li or Mg loading for Li/CaO
and Mg:Al hydrotalcite catalysts in the transesterication of glyceryl tributyrate with methanol.

and a drop in surface area and corresponding catalytic


activity. In another publication,43 we synthesised MgAl
hydrotalcite materials with different Mg compositions by
an alkali-free coprecipitation route. Trace alkali residues
found in materials prepared using conventional methods
using alkali carbonates and hydroxides can be problematic
due to leaching and product contamination in liquid phase
reactions. All materials showed very good performance for
the liquid phase transesterication of glyceryl tributyrate
with methanol for biodiesel production. The rate increases
steadily with Mg content (Fig. 1), with the Mg rich catalyst an order of magnitude more active than conventional
solid base MgO. The rate of reaction also correlates with
intralayer electron density, which can be associated with
increased basicity. Surface characterization reveals that in
both classes of solid base the electronic state of the surface
Li or Mg species can be correlated with catalyst activity.
Xie et al.44 tested alumina loaded with different potassium precursor catalysts for the transesterication of
soybean oil. The conversion to methyl esters over the catalysts is in the following order: KI/Al2 O3 > KF/Al2 O3 >
KOH/Al2 O3 > KNO3 /Al2 O3 > K2 CO3 /Al2 O3 > KBr/
Al2 O3 . KI/Al2 O3 demonstrated superior catalytic activity
compared to the other catalysts; however, to achieve the
highest conversion correspondingly high catalyst loadings (35 wt%) and activation temperatures are needed,
and none of these materials were tested for leaching or
reusability.

The most common acid catalysts employed are H2 SO4 and


HCl. Freedman et al.23 showed that the methanolysis of
soybean oil, in the presence of 1 mol% of H2 SO4 with
an alcohol/oil molar ratio of 30:1 at 65  C, takes 50 h to
reach complete conversion of the vegetable oil (>99%),
while butanolysis (at 117  C) and ethanolysis (at 78  C),
using the same quantities of catalyst and alcohol, take 3
and 18 h, respectively.
Al-Widyan et al.47 worked on the transesterication of
used palm oil under various conditions using different concentrations of HCl, H2 SO4 , and excess ethanol. They concluded that at higher catalyst concentrations (1.52.25 M)
J. Biobased Materials and Bioenergy 1, 112, 2007

Narasimharao et al.

2.2.2. Heterogeneous Catalysts


The use of commercial sulfonic ion-exchange resin was
also reported for production of biodiesel.48 The methanolysis of babassu and soybean oil was compared using
Amberlyst-15 with sulfuric acid as a catalyst, with the
cationic-exchange resin found to exhibit better activity
than the homogeneous catalyst.49 The methanolysis of soybean oil was carried out using different solid super acids
such as tungstated zirconia-alumina (WZA), sulfated tin
oxide (STO), and sulfated zirconia-alumina (SZA),50 of
which the WZA catalyst was the most effective, achieving conversions >90% at temperatures above 250  C after
20 h. Amberlyst-15, SZ, Naon NR50, and WZ showed
reasonably good activities at moderate temperatures of
60  C, indicating that these are suitable alternatives to the
homogeneous systems which can overcome the drawbacks
of corrosion and handling of liquid mineral acids.
Transesterication has also been conducted on a simple TG (triolein) with ethanol using various ion-exchange
resins as a heterogeneous catalyst.51 The anion-exchange
resins with a lower cross-linking density and a smaller
particle size gave a high reaction rate and high conversion. Another advantage of the resins is that they could
be recycled in batch transesterication without any loss in
the catalytic activity. The anion-exchange resins exhibited
much higher catalytic activities than their cation-exchange
counterparts. A continuous transesterication reaction was
carried out using an expanded bed reactor packed with
the most active resin. The reactor system permitted the
continuous production of ethyl oleate with a high conversion. Kaita et al.52 synthesized aluminum phosphate
catalysts with various Al/P molar ratios and used the resultant materials for the transesterication of kernel oil with
methanol. According to the authors, these catalysts were
thermally stable with good reactivity and selectivity to
methyl esters; however, their applications still required
high temperatures (200  C) and high methanol-to-oil molar
ratios (60:1).
2.3. Enzyme Catalysts
Although enzyme-catalyzed transesterication processes
are not yet commercially developed, new results have
been reported in recent articles and patents. The common
aspects of these studies involve optimization of the
J. Biobased Materials and Bioenergy 1, 112, 2007

reaction conditions (solvent, temperature, pH, type of


microorganism which generates the enzyme, etc.) in
order to establish suitable characteristics for an industrial
application.53 However, the reaction yields as well as the
reaction times are still unfavorable compared to those of
the base-catalyzed systems.
Lipase enzymes can also catalyze methanolysis of
triglycerides. The most promising results were obtained by
Fukuda et al. using immobilized Candida Antarctica
lipase (Novozym 435).54 Shimada et al.55 found that
Novozym435 was inactivated by shaking it in a mixture
containing more than 1.5 M eq. of methanol to oil. Above
this concentration, methanol is partially present as small
droplets in the oil phase. These droplets are believed to
cause enzyme deactivation. Therefore, methanol was
added stepwise; after the addition of the third methanol
equivalent, conversion to methyl esters was almost complete. The enzyme could be reused 50 times without loss
of activity. The occurrence of free fatty acids did not affect
the enzyme catalyst. Before the inlet of every reactor,
1 M eq. was added to the feed. Samukawa et al.56
reported an even more dramatic increase of the lipase efciency when it was pretreated by a consecutive incubation in methyl ester and oil prior to reaction. The use
of Novozym435 in methanolysis of triglycerides is also
reported in supercritical carbon dioxide at 24.1 MPa and
50  C.57 High yields (9095%) of fatty acid methyl esters
could be obtained when the reaction was carried out at
molar methanol/oil ratios of 25:1.
Hsu et al.58 studied the optimization of alkyl ester production from grease using a phyllosilicate solgel immobilized lipase. According to the studies, it was concluded
that the immobilized lipase was active from 40 to 70  C.
Ester contents of 6097% were highest when using a
ratio of reactants of 2 mmol grease to 8 mmol alcohol
and the biocatalyst was 10% (w/w) in the presence of a
molecular sieve. Watanabe et al.59 investigated the enzymatic conversion of waste edible oil to biodiesel in a
xed-bed bioreactor. Three-step methanoloysis and onestep methanoloysis of waste oil were conducted using Candida antarctica lipase. As a result, 90% conversion from
waste oil to biodiesel was obtained in both reactions.
In general enzymatic reaction selectivity is high, and
enzymes can be immobilized on support materials. However, enzymes are very expensive and exhibit unstable
activities with reaction rates much lower than those possible with homogeneous base catalysts.
2.4. Supercritical Transesterication
Saka and Kusdiana60 have developed a catalyst-free
method for biodiesel fuel production by employing supercritical methanol. The supercritical treatment at 350  C,
43 MPa, and 240 s with a molar ratio of 42:1 in methanol
is the optimum condition for transesterication of rapeseed
oil to biodiesel fuel. The great advantage of this method
7

REVIEW

biodiesel with a lower specic gravity was produced in


shorter reaction times. In their work, the variation of the
specic gravity of the end product with time was used as
an indicator for the effectiveness and completeness of the
conversion process. Lower values of specic gravity were
interpreted to indicate that more of the heavy glycerine
was removed, which, in turn, meant more complete reaction. H2 SO4 offered better conversion levels than HCl at a
catalyst concentration of 2.25 M.

Catalysts in Production of Biodiesel: A Review

REVIEW

Catalysts in Production of Biodiesel: A Review

was that free fatty acids present in the oil could be simultaneously esteried in the supercritical solvent.
Variables such as the molar ratio of alcohol to vegetable
oil and reaction temperature were investigated during
the transesterication within this supercritical media.61
Increasing the reaction temperature within the supercritical
regime resulted in increased ester conversion.
Diosakou et al.62 reported an 85% conversion of soybean oil into methyl esters after 10 h reaction at 235  C
and 6.2 MPa with a methanol:oil molar ratio of 21:1. The
rate of the reaction can be dramatically improved when
the reaction proceeds in supercritical methanol. Methanol
reaches its supercritical stage at 239  C and 8.1 MPa and is
believed to dissolve the oil completely. Under supercritical
conditions of 350  C and 45 MPa, rapeseed oil was completely esteried in 240 s with a methanol:oil molar ratio
of 42.63 Furthermore the presence of free fatty acids did
not alter the yield or rate of the supercritical methanolysis
reaction.
This new supercritical methanol process requires a
shorter reaction time and a simpler purication procedure
because of the absence of dissolved catalyst. Of course
this method does necessitate high temperature and pressure
(and therefore energy costs) and also requires an expensive
workup because of dissolution of the glycerol by-product
in methanol under these reaction conditions.
2.5. Esterication of Free Fatty Acids
The presence of free fatty acids in oils causes signicant
processing problems in standard biodiesel manufacturing,
since the free fatty acid is readily saponied by the homogeneous alkali catalyst used to transesterify triglycerides,
leading to a loss of catalyst as well as increased purication costs.64 The amount of free acids contained in different feedstocks is given in Table II. Turck et al.65 have
investigated the negative inuence of high FFA contents
on the base-catalyzed transesterication of triglycerides.
Free fatty acids react with the basic catalyst added for
the reaction and give rise to soap, as a result of which
some portion of the catalyst is neutralized and is therefore
no longer available for transesterication. These high FFA
content oils/fats are processed with an immiscible basic
glycerol phase so as to neutralize the free fatty acids and
cause them to pass over into the glycerol phase by means
of monovalent alcohols.
The main approach for improve the processing of free
fatty acid oils is to rst esterify the free fatty acids to
Table II.

Amount of free fatty acids in different feed stocks.

Feed stock
Rened vegetable oils
Crude soybean oil
Restaurant waste grease
Animal fat
Trap grease

Amount of free fatty acid (wt%)


<0.05%
0.30.7%
27%
530%
75100%

Narasimharao et al.

alkyl esters in the presence of an acidic catalyst before


transesterication. The pretreated oils, in which the free
fatty acid content is lowered to no more than 0.5 wt%,
can then be processed under standard transesterication
reaction conditions.66
2.5.1. Homogeneous Acid Catalysts
Van Gerpen et al.67 studied a technique to reduce the free
fatty acid content of restaurant waste and animal fats, using
an acid-catalyzed pretreatment to esterify the FFAs before
transesterifying the triglycerides with an alkaline catalyst
to complete the reaction. Initial process development was
performed with synthetic mixtures that contained between
20% and 40% FFAs, prepared by using palmitic acid. The
research showed that the acid level of the high-FFAs feedstocks could be reduced to <1% with a two-step, acidcatalyzed pretreatment reaction. The two-step pretreatment
reaction was performed with actual feedstocks with yellow
grease that contained 12% FFAs and brown grease that
contained 33% FFAs. After reducing the acid level, fuelgrade biodiesel was produced via an alkaline-catalyzed
transesterication reaction.
A new procedure for the preparation of methyl esters
from free fatty acids under mild conditions was
investigated.68 Free fatty acids are dissolved in a mixture
of chloroform-methanolic HCl-cupric acetate and kept at
room temperature for 30 min for complete esterication.
The method is also found to be suitable for esterifying
normal long-chain acids and very long chain (C12 C16 )
acids.
Simple Brnsted acid catalysts are the most frequently
used catalysts for the direct esterication. For simple reactions, e.g., with methanol or ethanol, an organic resin with
sulfonic acid groups is an industrially applied catalyst.69
However, esterication is generally slower when either
alcohol or acid is more sterically hindered, as in the case
of long-chain fatty alcohols or, e.g., the tertiary acid abietic
acid. In such cases, sulfuric acid still seems to be the catalyst of choice in industry, although HCl, dissolved arylsulfonic acids, H3 PO4 , polyphosphoric acid, and mixtures
of these catalysts are also employed. Even if esterication
is a very old reaction, there is still an intensive search for
newer catalysts which target the following goals.70 First,
the reaction should be run with equimolar amounts of acid
and alcohol, instead of an excess of one or both. Next, it
is desirable to perform the reaction at low temperatures
and in mild, almost neutral conditions, so that the reaction
can proceed even in the presence of acid-labile groups.
Finally, a catalyst should be removable from the products
and reusable.
Lewis acids may be preferred over Brnsted acids for
several reasons, most notably to avoid alcohol dehydration
or racemization by the Brnsted acid or to create conditions that are compatible with acid-labile groups. However, the distinction between Lewis and Brnsted acids is
J. Biobased Materials and Bioenergy 1, 112, 2007

Narasimharao et al.

often difcult in a reaction producing water, such as direct


esterication. For instance, Lewis acid compounds such as
BF3 or zinc oxalate form protonic acids in contact with
water or an alcohol.71

The use of liquid inorganic or mineral acid catalysts in


esterication reactions requires neutralization and separation steps to be incorporated into the process to isolate the
product. Therefore, it would be desirable to perform the
esterication pretreatment step with a heterogeneous acidic
catalyst, which would signicantly simplify the separation
and thus reduce the net number of processing steps.72
Several zeolites, such as modied H-Y, H-Beta, and HZSM-5, have been employed as esterication catalysts. In
order to expel the water formed, hydrophobicity of the
zeolite lattice is expected to be an important parameter.
Which zeolite is eventually most suitable for a specic
reaction depends on the polarity and miscibility of the
acid and alcohol reactants.73 Mesoporous ordered materials of the MCM family are generally not sufciently acidic
to catalyze esterication, but modication with sulfonic
acid groups has resulted in stable and active materials.74 75
Heteropolyacids like H3 PW12 O40 and H4 SiW12 O40 that
have been supported on carbon76 have also been used
in esterication; however, care must be taken to avoid
leaching, as H3 PW12 O40 exhibits high solubility in polar
media. Ion exchanged heteropoly compounds such as
Cs25 H05 PW12 O40 are heterogeneous water tolerant catalysts which are more promising for use in esterication
reactions.77
There are a few reports of the synthesis of fatty
acid esters over solid acid catalysts including sulfated
metal oxides78 79 and heteropoly acids.80 The use of
ion exchange resins was reported by Steinigeweg and
Gmehling81 for the synthesis of methyl decanoate. However, such resin catalysts could not be used for esterications conducted at more than 120  C, because of their
limited thermal stability. Omota et al.82 used sulfated zirconia as the catalyst for esterication of dodecanoic acid
with 2-ethylhexanol. Zinc acetate supported over functionalized silica gel was reported to have good characteristics in terms of its surface area, pore size, and thermal
stability and gave high conversion for the esterication
reaction.83 Aafaqi et al.84 studied the synthesis of isopropyl
palmitic ester utilizing zinc acetate supported over silica
gel as a catalyst in a batch reactor. Organosulfonic acidfunctionalized mesoporous silicas have been demonstrated
to have activity comparable to that of H2 SO4 in the conversion of fatty acids to methyl esters.85 Recently, we have
reported on a mesoporous tungstated zirconium phosphate
solid acid catalyst which has high activity in the esterication of palmitic acid.86 Optimum activity was observed
for 6.37.8 wt% W loadings, where we believe highly dispersed Keggin-like clusters are formed.
J. Biobased Materials and Bioenergy 1, 112, 2007

2.6. Simultaneous Transesterication and


Esterication
Canakci et al.67 employed an integrated esterication and
transesterication method, wherein the high-FFA feedstock was initially treated using an acidic catalyst (H2 SO4 )
to reduce the FFA level to less than 1%. This pretreated
feedstock was then transesteried with methanol, using
an alkaline catalyst (KOH). The two-step, acid-catalyzed
esterication followed by an alkaline-catalyzed reaction
improved the methyl ester yield from waste cooking oil,
and the reaction rate increased with the amount of acid
catalyst.
Di Serio et al.87 reported a method for the simultaneous esterication and transesterication of waste oils using
homogeneous Lewis acids based on carboxylic salts of the
metals Cd, Mn, Pb, and Zn. Catalytic activities for the
transesterication of waste oil (at a molar ratio of oil to
alcohol of 1:12 and a temperature of 200  C for 200 min)
by these catalysts was related to the Lewis acid strength of
the metals and to the structure of the anion. The authors
also studied the inuence of water and FFA content in
oils on the catalyst activity. They found the water has a
strong depressive effect on the activity of all the investigated metal salts. This effect can be attributed to the
interaction of water with the cation of the catalysts, which
decreases their acid strength. The presence of FFA and
water formed from the esterication reaction is, however,
detrimental to conversion of triglyceride. Stearates showed
better performance than acetates, because of their greater
solubility in the oil.
Lepper and Friesenhagen88 have patented a two-stage,
acidic and basic catalytic process for the production of
fatty acid esters of short-chain aliphatic alcohols from fats
and/or oils with an FFA content of >1%. However, these
homogeneous acid and base catalysts suffer serious drawbacks associated with catalyst isolation from the product
and leaching as described in earlier sections.
2.7. Biodiesel from Vegetable Oil via Pyrolysis
2.7.1. Noncatalytic Pyrolysis
The pyrolysis of fats has been investigated for more than
100 years, especially in areas of the world without
petroleum deposits. Pyrolysis is the thermal degradation
of vegetable oils by heat in the absence of oxygen,
which results in the production of alkanes, alkenes, alkadienes, carboxylic acids, aromatics, and small amounts of
gaseous products.89 The mechanism of pyrolysis of triglycerides was given by Schwab et al.90 The same authors
reported the decomposition and distillation of91 soybean
oil thermally in air and in a standard ASTM distillation apparatus. They also used Safower oil, which
possesses a high oleic oil control. The total identied hydrocarbons obtained from the distillation of soybean and high oleic Safower oils were 7377% and
9

REVIEW

2.5.2. Heterogeneous Acid Catalysts

Catalysts in Production of Biodiesel: A Review

REVIEW

Catalysts in Production of Biodiesel: A Review

8088%, respectively. The main components were alkanes


and alkenes, which accounted for approximately 60% of
the total weight. Carboxylic acids accounted for another
9.616.1%.
Rapeseed oil was pyrolyzed to produce a mixture
of methyl esters in a tubular reactor between 500 and
850  C under nitrogen.92 The conversion of methyl colzate
increased with an increase of the temperature of pyrolysis. To illustrate the distribution of cracking products as a
function of pyrolysis temperature, the selectivities of products (hydrocarbons, CO, CO2 , and H2 ) were determined
between 550 and 850  C for a constant residence time
of 320 min. The principal products were linear 1-olens,
n-parafn, and unsaturated methyl esters. High temperatures gave high yields of light hydrocarbons (66% molar
ratio at 850  C). However, the equipment for thermal
cracking and pyrolysis is expensive for modest throughputs, and furthermore while the products are chemically
similar to petroleum-derived gasoline and diesel fuel, the
absence of oxygen during thermal processing also removes
any of the environmental benets obtained from using an
oxygenated fuel. Pyrolysis therefore produces some low
value materials and, sometimes, more gasoline than diesel
fuel.2
2.7.2. Catalytic Pyrolysis
Catalytic cracking of vegetable oils to produce liquid biofuels has been studied by Pioch et al.93 This work was
performed in a xed bed reactor with a commercial nickelbased catalyst for naphtha steam reforming. Sunower oil
was completely converted to hydrogen, methane, and carbon oxides, except for runs performed at the lowest temperatures. The hydrogen yield ranged from 72% to 87%
of the stoichiometric potential, depending on the steamto-carbon ratio and the temperature at the catalyst bed,
which governed the equilibrium among the gas species.
Copra oil and palm oil were also tested for pyrolysis over
SiO2 /Al2 O3 at 450  C to produce biodiesel fuels. The
chemical compositions of diesel fractions were similar to
those of fossil fuels.

3. FUTURE SCOPE OF RESEARCH


Biodiesel has become increasingly attractive because of its
environmental benets and the fact that it is made from
renewable resources. The remaining challenges are its cost
and limited availability of fat and oil resources. There are
two aspects of the cost of biodiesel, the costs of raw material (fats and oils) and the cost of processing. The cost
of raw materials accounts for 6075% of the total cost
of biodiesel fuel. The use of spent cooking oil can lower
the cost signicantly; however, the quality of used cooking
oils can be low and variable. Studies are therefore needed
to nd a cheaper way to utilize waste cooking oils to make
biodiesel fuel. There are several choices: the rst choice
10

Narasimharao et al.

is removing free fatty acids from the cooking oil before


transesterication, using acid-catalyzed esterication; the
second one is transesterication at high temperature and
pressure (supercritical).
In terms of production cost, there also are two aspects,
the transesterication process and by-product (glycerol)
recovery. A continuous transesterication process using
xed bed reactors would be one means to lower the production cost by simplifying product separation. The foundations of this process are a shorter reaction time and
greater production capacity. The recovery of high-quality
glycerol with less alkali and water contamination due to
elimination of aqueous quench steps would also improve
production costs.
Biodiesel properties are strongly inuenced by the properties of the individual fatty esters. It therefore appears
reasonable to enrich the fuel with certain fatty esters with
desirable properties in the fuel in order to improve the
properties of the whole fuel. It may be possible in the
future to improve the properties of biodiesel by means of
genetic engineering of the parent oils, which could eventually lead to a fuel enriched with certain fatty acids, possibly oleic acid, that exhibit a combination of improved fuel
properties.
Depending on the water and FFA content in the feedstock, a transesterication method should be selected. If
the FFA and water contents are <1 wt% and <0.5 wt%,
respectively, then a base catalyst is more suitable for
the ester production. If the FFA content of oil is high
(>1 wt%), then an acid catalyst is a better choice.
However, because of the requirement of high catalyst concentration and high molar ratio and because of corrosion
problems, homogeneous catalysts are also not recommended for the transesterication of waste cooking oil.
Such a two-step homogeneous method (an acid-catalyzed
step, followed by an alkaline-catalyzed step) would make
biodiesel production more costly because of the extra processing steps required. Enzyme-catalyzed transesterication is a very good alternative to all chemical-catalyzed
reactions but requires extensive research prior to commercialization. The catalyst-free supercritical methanol
method has great potential for biodiesel production from
waste cooking oil; however, the requirements of high temperature (350  C), high pressure (45 MPa), and high molar
ratio of oil to alcohol (1:42) make the use of this process difcult on an industrial scale. It is recommended
that dual bed catalytic reactors, using a combination of
solid acid and base catalysts, are developed to minimize
the number of reaction steps and/or bifunctional (acid
base) heterogeneous catalyst are developed which can catalyze both esterication and transesterication reactions in
a single pot.
Acknowledgments: K. N. Rao thanks the Royal Society, London, UK, for the award of a Royal Society International Incoming Fellowship from India.
J. Biobased Materials and Bioenergy 1, 112, 2007

Narasimharao et al.

References

J. Biobased Materials and Bioenergy 1, 112, 2007

36. U. Schuchardt, R. M. Vargas, and G. Gelbard, J. Mol. Catal. 109,


37 (1996).
37. S. K. F. Peter, R. Ganswindt, H. P. Neuner, and E. Weidner, Eur. J.
Lipid Sci. Technol. 104, 324 (2002).
38. G. J. Suppes, K. Bockwinkel, S. Lucas, J. B. Botts, M. H. Mason,
and A. J. Heppert, J. Am. Oil Chem. Soc. 78, 139 (2001).
39. E. Leclercq, A. Finiels, and C. Moreau, J. Am. Oil Chem. Soc. 78,
1161 (2001).
40. G. J. Suppes, A. D. Mohanprasad, E. J. Doskocil, J. M. Pratik, and
M. J. Goff, Appl. Catal. A: Gen. 257, 213 (2004).
41. H.-J. Kim, B.-S. Kang, M.-J. Kim, Y. M. Park, D.-K. Kim, J. S. Lee,
and K.-Y. Lee, Catal. Today 93, 315 (2004).
42. R. S. Watkins, A. F. Lee, and K. Wilson, Green Chem. 6, 335 (2004).
43. D. G. Cantrell, L. J. Gillie, A. F. Lee, and K. Wilson, Appl. Catal.
A: Gen. 287, 183 (2005).
44. W. Xie and H. Li, J. Mol. Catal. A: Chem. 255, 1 (2006).
45. K. J. Harrington and C. D. Arcy-Evans, Ind. Eng. Chem. Prod. Res.
Dev. 24, 314 (1985).
46. W. Stoffel, F. Chu, and E. H. Ahrens, Jr., Anal. Chem. 31, 307
(1959).
47. M. I. Al-Widyan and A. O. Al-Shyoukh, Bioresour. Technol. 85, 253
(2002).
48. S. Abro, Y. Pouilloux, and J. Barrault, Stud. Surf. Sci. Catal. 108,
539 (1997).
49. G. Vicente, A. Coteron, M. Martinez, and J. Aracil, Ind. Crops Products 8, 29 (1998).
50. S. Furuta, H. Matsuhashi, and K. Arata, Catal. Commun. 5, 721
(2004).
51. D. E. Lpez, J. G. Goodwin, Jr., D. A. Bruce, and E. Lotero, Appl.
Catal. A: Gen. 295, 97 (2005).
52. J. Kaita, T. Mimura, N. Fukuoda, and Y. Hattori, Catalysts for transesterication. U.S. Patent 6407269 (2002).
53. U. Schuchardta, R. Serchelia, and R. M. Vargas, J. Braz. Chem. Soc.
9, 199 (1998).
54. H. Fuduka, A. Kondo, and H. Noda, J. Biosci. Bioeng. 92, 405
(2001).
55. Y. Shimada, Y. Watanabe, T. Samukawa, A. Sugihara, H. Noda,
H. Fukuda, and Y. Tominaga, J. Am. Oil Chem. Soc. 76, 789 (1999).
56. T. Samukawa, M. Kaieda, T. Matsumoto, K. Ban, A. Kondo,
Y. Shimada, H. Noda, and H. Fukuda, J. Biosci. Bioeng. 90, 180
(2000).
57. M. A. Jackson and J. W. King, J. Am. Oil Chem. Soc. 73, 353 (1996).
58. A. Hsu, K. Jones, and W. N. Marmer, J. Am. Oil Chem. Soc. 78,
585 (2001).
59. Y. Watanabe, Y. Shimada, A. Sugihara, and Y. Tominaga, J. Am. Oil
Chem. Soc. 78, 703 (2001).
60. S. Saka and K. Dadan, Fuel 80, 225 (2001).
61. D. Kusdiana and S. Saka, Bioresour. Technol. 91, 289 (2004).
62. M. Diasakov, A. Loulodi, and N. Papayannakos, Fuel 77, 1297
(1998).
63. S. Saka and D. Kusdiana, Fuel 80, 225 (2001).
64. M. Canakci and J. V. Gerpen, Trans. ASAE 42, 1203 (1999).
65. R. Turck, Method for producing fatty acid esters of monovalent alkyl
alcohols and use thereof. US Patent 0156305 (2002).
66. M. Canakci and J. V. Gerpen, Trans. ASAE 44, 1429 (2001).
67. M. Canakci and J. V. Gerpen, A pilot plant to produce biodiesel from
high free fatty acid feed stocks, ASAE paper No. 01-6049, ASAE,
St. Joseph, MI, USA (1999).
68. M. Hoshi, M. Williams, and Y. Kishimoto, J. Lipid Res. 14, 599
(1973).
69. Ullmanns Encyclopedia of Technical Chemistry, 6th edn., Electronic
Release, Wiley-VCH (2000); Baileys Industrial Oil & Fat Products,
edited by Y. H. Hui, Wiley-Interscience, New York (1996).
70. J. Otera, Angew. Chem. 113, 2099 (2001).
71. P. K. Kadaba, Synthesis 628 (1972).
72. G. R. Peterson and W. P. Sacarrah, J. Am. Oil Chem. Soc. 61, 1593
(1984).

11

REVIEW

1. S. J. Clark, L. Wagner, M. D. Schrock, and P. G. Pinnaar, J. Am. Oil


Chem. Soc. 61, 1632 (1984).
2. F. Ma and M. Hanna, Bioresour. Technol. 70, 1 (1999).
3. S. Romano, Vegetable oils: A new alternative. Vegetable Oils Fuels:
Proc. Int. Conf. on Plant and Vegetable Oils as Fuels, ASAE,
St. Joseph, MI, USA (1982), pp.
4. M. S. Graboski and R. L. Mc Cormick, Prog. Energy Combust. Sci.
24, 125 (1998).
5. E. G. Shay, Biomass Bioenergy 4, 227 (1993).
6. W. R. Nitschke, C. M. Wilson, and R. Diesel, Pioneer of the Age of
Power, The University of Oklahoma Press, USA (1965).
7. D. Y. Z. Chang, J. H. Van Gerpen, I. Lee, L. A. Johnson, E. G.
Hammond, and S. J. Marley, J. Am. Oil Chem. Soc. 73, 1549 (1996).
8. H. Fukuda, A. Kondo, and H. J. Noda, Biosci. Bioeng. 92, 405
(2001).
9. Mag. Stephan Friedrich, World wide review of the commercial
production of biodieselA technological, economic and ecological
investigation based on case studies, http://itpn.wu-wien.ac.at.
10. F. Eibensteiner and H. Danner, Biodiesel in Europe, System Analysis, Non-Technical-Barriers, Wels (2000), p. 32.
11. M. Canakcia, A. Erdilb, and E. Arcakliogluc, Appl. Energy 83, 594
(2006).
12. A. Srivastava and R. Prasad, Renewable Sustainable Energy Rev. 4,
111 (2000).
13. G. Knothe, Fuel Processing Technol. 86, 1059 (2005).
14. K. S. Tyson, Biodiesel, Handling and Use Guidelines. NREL,
Golden CO (2001).
15. P. Talley, Biodiesel. Render (2004).
16. http://www.biodiesel.co.uk
17. C. E. Goering and B. Fry, J. Am. Oil Chem. Soc. 61, 1627 (1984).
18. J. W. Alencar, P. B. Alves, and A. A. Craveiro, J. Agric. Food Chem.
31, 1268 (1983).
19. R. Altin, S. Cetinkaya, and H. S. Yucesu, Energy Convers. Management 42, 529 (2001).
20. M. W. Formo, Physical properties of fats and fatty acids. Baileys
Industrial Oil and Fat Products, 4th edn., John Wiley and Sons,
New York (1979), Vol. 1.
21. H. Noureddini and D. Zhu, J. Am. Oil Chem. Soc. 74, 1457 (1997).
22. B. Freedman, E. H. Pryde, and T. L. Mounts, J. Am. Oil Chem. Soc.
61, 1638 (1984).
23. B. Freedman, R. O. Buttereld, and E. H. Pryde, J. Am. Oil Chem.
Soc. 63, 1375 (1986).
24. D. Darnoko and M. Cheryan, J. Am. Oil Chem. Soc. 77, 1263 (2000).
25. R. M. Vargas, D. Sc. Thesis, Universidade Estadual de Campinas,
Campinas, Brazil (1996).
26. U. Schuchardt, R. M. Vargas, and G. Gelbard, J. Mol. Catal. 99, 65
(1995).
27. F. R. Abreu, D. G. Lima, E. H. Ham, S. Einloft, J. C. Rubim, and
P. A. Z. Suarez, J. Am. Oil Chem. Soc. 80, 601 (2003).
28. F. R. Abreu, D. G. Lima, E. H. Ham, C. Wolf, and P. A. Z. Suarez,
J. Mol. Catal. A: Chem. 209, 29 (2004).
29. T. Cercce, S. Peter, and E. Weidner, Ind. Eng. Chem. Res. 44, 9535
(2005).
30. E. Lotero, Y. Liu, D. E. Lopez, K. Suwannakarn, D. A. Bruce, and
J. G. Goodwin, Jr., Ind. Eng. Chem. Res. 44, 5353 (2005).
31. H. E. Hoydonckx, D. E. De Vos, S. A. Chavan, and P. A. Jacobs,
Top. Catal. 27, 83 (2004).
32. A. Corma, S. Iborra, S. Miquel, and J. Primo, J. Catal. 173, 315
(1998).
33. A. Corma, S. B. A. Hamid, S. Iborra, and A. Velty, J. Catal. 234,
340 (2005).
34. U. Schuchardt and O. C. Lopes, Chem. Abstr. 101, P93246 (1984).
35. R. Sercheli, A. L. B. Ferreira, M. C. Guerreiro, R. M. Vargas, R. A.
Sheldon, and U. Schuchardt, Tetrahedron Lett. 38, 1325 (1997).

Catalysts in Production of Biodiesel: A Review

REVIEW

Catalysts in Production of Biodiesel: A Review

Narasimharao et al.

73. N. Sanchez, M. Martinez, J. Aracil, and A. Corma, J. Am. Oil Chem.


Soc. 69, 1150 (1992).
74. W. D. Bossaert, D. E. De Vos, W. Van Rhijn, J. Bullen, P. J. Grobet,
and P. A. Jacobs, J. Catal. 182, 156 (1999).
75. K. Wilson, A. F. Lee, D. J. Macquarrie, and J. H. Clark, Appl. Catal.
A: Gen. 228, 127 (2002).
76. Y. Izumi and K. Urabe, Chem. Lett. 663 (1981).
77. T. Okuhara, Chem. Rev. 102, 3641 (2002).
78. S. Furuta, H. Matsuhashi, and K. Arata, Appl. Catal. A: Gen. 269,
187 (2004).
79. S. Ramu, N. Lingaiah, B. L. A. Prabhavathi Devi, R. B. N. Prasad,
I. Suryanarayana, and P. S. Sai Prasad, Appl. Catal. A: Gen. 276,
163 (2004).
80. B. Y. Giri, K. Narasimha Rao, B. L. A. Prabhavathi Devi,
N. Lingaiah, I. Suryanarayana, R. B. N. Prasad, and P. S. Sai Prasad,
Catal. Commun. 6, 788 (2005).
81. S. Steinigeweg and J. Gmehling, Ind. Eng. Chem. Res. 42, 3612
(2003).
82. F. Omota, A. C. Dimian, and A. Bliek, Chem. Eng. Sci. 58, 3175
(2003).
83. R. Nava, T. Halachev, R. Rodriguez, and V. M. Castano, Microporous
Mesoporous Mater. 78, 91 (2005).

84. R. Aafaqi, A. R. Mohamed, and S. Bhatia, J. Chem. Technol.


Biotechnol. 79, 1127 (2004).
85. I. K. Mbaraka and B. H. Shanks, J. Catal. 229, 365 (2005).
86. K. N. Rao, A. Sridhar, A. F. Lee, S. J. Tavener, N. A. Young, and
K. Wilson, Green Chem. 790, 8 (2006).
87. M. Di Serio, R. Tesser, M. Dimiccoli, F. Cammarota, M. Nasatasi,
and E. Santacesaria, J. Mol. Catal. A: Chem. 239, 111 (2005).
88. H. Lepper and L. Friesenhagen, Process for the production
of fatty acid esters of short-chain aliphatic alcohols from fats
and/or oils containing free fatty acids. U.S. Patent 4,608,202
(1986).
89. N. O. V. Sonntag, Reactions of fats and fatty acids. 4th edn., Baileys
Industrial Oil and Fat Products, edited by D. Swern, John Wiley &
Sons, New York (1979), Vol. 1.
90. A. W. Schwab, G. J. Dykstra, E. Selke, S. C. Sorenson, and E. H.
Pryde, J. Am. Oil Chem. Soc. 65, 1781 (1988).
91. A. W. Schwab, M. O. Bagby, and B. Freedman, Fuel 66, 1372
(1987).
92. F. Billaud, V. Dominguez, P. Broutin, and C. Busson, J. Am. Oil
Chem. Soc. 72, 1149 (1995).
93. D. Pioch, P. Lozano, M. C. Rasoanantoandro, J. Graille, P. Geneste,
and A. Guida, Oleagineux 48, 289 (1993).

Received: 24 July 2006. Revised/Accepted: 29 September 2006.

12

J. Biobased Materials and Bioenergy 1, 112, 2007

Anda mungkin juga menyukai