You are on page 1of 10

Journal of Industrial and Engineering Chemistry 21 (2015) 110

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Review

An overview of the role of ionic liquids in biodiesel reactions


Nawshad Muhammad a,*, Yasir A. Elsheikh b, Muhammad Ibrahim Abdul Mutalib c,
Aqeel Ahmed Bazmi d, Rahmat Ali Khan e, Hidayatullah Khan e, Sikander Raq d,
Zakaria Man c, Ihsnullah khan c
a

Interdisciplinary Research Center for Biomedical Materials (IRCBM), COMSATS Institute of Information Technology,
Lahore, Pakistan
Department of Chemical and Materials Engineering, Faculty of Engineering, King Abdulaziz University, P.O. Box 344, Rabigh 21911, Saudi Arabia
c
Department of Chemical Engineering, Universiti Teknologi PETRONAS (UTP), Bandar Seri Iskandar, 31750 Tronoh, Perak, Malaysia
d
Department of Chemical Engineering, COMSATS Institute of Information Technology, Lahore, Pakistan
e
Department of Biotechnology, Faculty of Biological Sciences, University of Science and Technology, Bannu, Khyber Pakhtunkhwa, Pakistan
b

A R T I C L E I N F O

Article history:
Received 2 July 2013
Accepted 17 January 2014
Available online 6 February 2014
Keywords:
Biodiesel
Transesterication
Ionic liquid
Catalyst
Solvent
Enzyme

A B S T R A C T

The concerns on the depleting petroleum resources and increasing environmental problems have driven
the scientic community worldwide to develop large-scale non-petroleum-based alternative fuels, such
as bioethanol and biodiesel. Biodiesel produced through the transesterication of vegetable oils or
animal fats are highly attractive. On the other hand, ionic liquids which possess properties that are more
environmental friendly have found signicant applications as solvents and catalysts for reaction and
separation. It is also beginning to nd its way in many of the chemical process applications and has
attracted signicant attention including biodiesel production. This paper provides a brief overview on
the feasibility of applying ionic liquids in biodiesel production for the purpose of powering diesel engines
for transportation and utility generation. The potential of applying ionic liquids as catalyst and solvent
for enzymatic production of biodiesel from feedstock is particularly highlighted.
2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Contents
1.
2.
3.

4.
5.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Biodiesel fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Role of ionic liquid as a catalyst for preparation of biodiesel . . . . . . . . . . . . . .
3.1.
Ionic liquid coupled with inorganic supporting material as catalyst for
3.2.
Ionic liquid solely as a catalyst for biodiesel reaction. . . . . . . . . . . . . . .
3.3.
Ionic liquid as deep eutectic solvents (DES) for biodiesel reaction . . . .
Role of ionic liquid as a solvent for biocatalyzed transesterication process . .
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
The need for renewable energy sources is becoming increasingly more important as the exploitation of petroleum resources

* Corresponding author.
E-mail addresses: nawshadchemist@yahoo.com,
nawshadmuhammad@ciitlahore.edu.pk (N. Muhammad).

...............
...............
...............
biodiesel reaction
...............
...............
...............
...............
...............
...............

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

1
2
3
4
5
6
6
8
9
9

increases over the years while its reserve declining at an alarming


rate worldwide. Biodiesel has been known as one of the potential
fuel substitutes derived from relatively cheap renewable biological
sources (vegetable oils or animal fats) which can be used widely for
powering diesel engines and utility systems [1]. Besides that, other
biological source such as oleaginous microbial biomass has also
been considered [2]. Earlier, the transesterication of unrened
oils has been performed by various researchers using catalysts

1226-086X/$ see front matter 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jiec.2014.01.046

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

such as sulfuric acid, hydrochloric acid, phosphoric acid, and


organic sulfonic acid [3,4] which are considered as hazardous and
non-green material.
As the pressure on using greener and less hazardous material
mounting up, several more environmental benign materials and
processes have been investigated. Among them is the use of
enzymes especially lipase for transesterication of vegetable oils
which have been signicantly studied over the past decade or so.
However, the developed processes such as the lipase-assisted
methanolysis of vegetable oils to produce fatty acid methyl esters
yielded poor conversion due to the well-known effect of lipase
deactivation by methanol [5].
Recently ionic liquids (ILs) have been found to be a very useful
medium that could assist and promote better reaction and
separation for various type and combination of enzymes used
for the biodiesel synthesis. The ILs, which is a new class of material,
are relatively non-volatile compared to the organic solvents and
their development has evolved tremendously over the years to
become increasingly more suitable media for many enzymatic
reactions [6]. ILs are considered to be unconventional but ecofriendly media widely considered in many emerging research area
particularly during recent times [7].
Besides being non-volatile, the ILs have excellent chemical and
thermal solubility which is also tuneable based on the combination
of cation and anion used. For the biological reactions such as the
biodiesel synthesis, the ILs are capable of dissolving a wide array of
substrates and more importantly, increasing their stability over a
longer period during the reaction [710]. Moreover, some of these
properties can be ne-tuned by changing the cation or anion of the
ionic liquids which has led to the ILs being termed as designer
solvent. The potential of IL as green solvents for biochemical
conversion has been well-known [7]. The ionic liquids (ILs) offer an
excellent media for many lipase-catalyzed esterication/transesterication reactions compared to the organic solvents wherein an
increased in the activity and specicity of the lipase enzyme was
observed [1115]. Many recent studies [1,1639] proved the ILs as
more environmental friendly solvents and catalysts for biodiesel
reactions and separation.
This study aims at reviewing the roles of ionic liquid in biodiesel
production reaction. The effect of varying the cation and anion
combinations and the acidity/basicity of the ionic liquid used as
well as the reaction conditions will be discussed. In addition, the
synergistic effects of the enzyme-based catalysis in ionic liquid will
also be considered.
2. Biodiesel fuel
The quickly growing demand for energy which has led to higher
depletion rate of fossil fuels reserve, coupled with the increasing
global warming threat due to increase in greenhouse gasses
emission have drawn signicant attention globally. This has led to
the impeccable need of nding renewable energy sources to ensure
future sustainability [1,2]. Biodiesel fuel i.e., fatty acid methyl
esters (FAMEs), is a renewable and environmentally friendly fuel.
There are different varieties of processes and feedstocks from
which the biodiesel can be synthesized such as castor beans, palm,
palm kernel, olive, sesame, corn, linseed, soybean, canola,
sunower, peanuts, coconut, oilseed radish, cotton, sunower,
babassu, algae, jatropha, sea mango, polanga, pongamia, animal fat
(butter, lard, tallow, grease and sh oil, etc.) and waste cooking oil
[6].
The current major challenge in making the biodiesel production
economically viable is the high price of the vegetable oils
compared to the fossil based diesel fuel. As a result, in some
countries, non-edible oils such as jatropha or waste cooking oil
have been used as a feedstock for biodiesel production [40].

Exploring ways to reduce the high cost of biodiesel is of much


interest in recent biodiesel research, especially for methods
focusing on minimizing the raw material cost. Comparing the
prices of biodiesel from the three different sources, palm biodiesel
shows the cheapest price; this indicates that palm biodiesel has
better economic potential than the other indicated oils. Among the
four leading vegetable oils traded on the world market, CPO is
much cheaper than canola, rapeseed or soybean [35]. In October
2008, the price of CPO was approximately half (55%) of the price of
crude soybean oil [16] (Fig. 1).
Zhang et al. [1], and Leung and Guo [26] reported that the cost of
raw materials accounted approximately between 70 and 95% of the
total cost of biodiesel production. Haas et al. [33] estimated the
cost of feedstock to be 88% of total biodiesel production cost for
soybean oil. Thus other feedstocks options are being evaluated by
several researchers, instead of the highly rened oil. Among others
include crude vegetable oil and recycled waste feedstock as the
acceptable ways to lower the biodiesel production cost [28,34].
The use of waste cooking oil instead of the virgin oil to produce
biodiesel is an option also considered for reducing the raw material
cost in view of its extensive availability besides signicantly
alleviating the problems associated with its waste handlings.
Although there were reports on some of the waste cooking oil
being used for soap preparation and as additive oil for fodder
making, but most of it end up in landlls and rivers causing
signicant environmental pollution [41]. This has led most of the
developed countries to set stringent policies on the safe disposal of
waste oil through the drainage system.
In 2002, the EU has enforced a ban on the utilization of all waste
oils as domestic animal feed which was seen as one of the effective
method of disposing waste cooking oil then. This is because during
the frying process, there were many harmful compounds formed
and feeding the waste cooking oil to the animal would eventually
cause it to enter human system through the food chain. This is
made worse especially if the waste cooking oil is reused several
times due to economic reasons. Under normal frying conditions,
the oil is boiled under atmospheric condition at temperature of
160190 8C for a relatively long period of time. Under such
condition, the waste oil will undergo various changes in its
physical and chemical properties. Generally the changes observed
in the oil comprised of the following; (i) increase in thickness and
viscosity, (ii) increase in specic heat, (iii) increase in surface
tension, and (iv) change toward a darker color. This could be
attributed to the increased presence of higher longer chain
saturated fats components in the oil as well as the other
components. There are three types of reactions taking place
during frying and they are mainly thermolytic, oxidative and
hydrolytic. These three reactions will continuously cause the

[(Fig._1)TD$IG]

Fig. 1. Prices of diesel fuel and biodiesel from different sources [36].

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

formation of many undesired and unsafe compounds if the oil is


used continually [42].
The amount of waste cooking oil generated in each country
varies depending on the use of vegetable oil by households and the
style of cooking adopted by the people. There is a signicant
variation in the cooking style between the Asians and the
Europeans. Generally, the waste cooking oils can be an effective
feedstock for biodiesel production as it will not be seen as
competing with the food source material in the case where virgin
oil is used. Although the waste cooking oil is available locally and in
abundance, the major issue is associated with the massive logistic
arrangement required for collecting the waste oil as the source
location is widespread within the locality in order to generate
sufcient quantities. The production of biodiesel from waste
cooking oil is one of the better ways to utilize it efciently and costeffectively.
The general synthetic route for biodiesel production is the
transesterication of vegetable oils (or animal fats) with alcohol
which is normally either methanol or ethanol. This reaction can be
catalyzed by acids, alkaline metal hydroxides, alkoxides, and nonionic bases (such as amines and amidines) [43]. However, these
methods have several drawbacks:
(i) The acid/base processes are often related to decay and
emulsication problems
(ii) The acid-catalyzed reactions are usually much slower than the
base-catalyzed processes [44,45]. However, the base-catalyzed process may cause unnecessary saponication of the
fatty acids feed [46].
(iii) A signicantly large excess of alcohol is required to drive the
equilibrium to the ester formation and to achieve the facile
separation of biodiesel from the produced glycerol by the
reaction hence creating recycling problem i.e., the investment
and operating cost required to recover the alcohol used.
(iv) Practically, the oils and fats are not soluble in alcohols
resulting in barriers for the triglyceride conversions.
(v) Other issues such as energy intensive, requirement of alkaline
waste-water treatment, and interference of free fatty acids and
water [47].

Therefore, a number of new approaches have been vigorously


pursued to circumvent these problems, such as the development
of heterogeneous catalysts [48,49] conducting the alcoholysis in
supercritical methanol [50], use of ILs-catalyzed transesterication [51,52] and use of the lipase-catalyzed transesterication
[53,54].
For product separation, there were several methods used
to separate FAMEs i.e., the biodiesel produced, from the
other components. However, there are only two generally
accepted methods to purify the biodiesel namely wet and dry
washing.
The more traditional wet washing method is widely used to
remove excess contaminants and available production chemicals
from the biodiesel. Since both glycerin and methanol are highly
soluble in water, water washing is very effective in removing both
contaminants and until recently was the most common method of
purication. It also has the advantage of removing any residual
sodium salts and soaps, the latter being a byproduct of high free
fatty acids (FFA) feeds, due to their water solubility. However, the
inclusion of additional water to the process offers many
disadvantages, including increased operating cost and production
time. A highly polluting liquid efuent is also generated thus
requiring special waste treatment. Signicant product losses was
also seen due to their carried over in the water phase. Furthermore,
emulsion formation was also observed during the processing of

waste cooking oils or other feeds that contain high FFA content due
to soap formation [55,56].
Dry washing has replaced the water washing where an ion
exchange resin or a magnesium silicate powder was used to
neutralize the impurities. Both of the dry washing methods are
being used in industrial plants currently. Other possible methods
that can be used to purify the produced biodiesel comprised of
membrane reactors and the addition of lime and phosphoric acid
[5759]. Nonetheless, there were numerous problems reported
especially on the costs and the complications faced in operating the
process at industrial scale.
3. Role of ionic liquid as a catalyst for preparation of biodiesel
ILs have been widely accepted as a new green chemical that is
capable of revolutionizing chemical processes due to their
interesting properties. The subject has created tremendous
interest within both the academia and the chemical industries.
ILs is basically a molten salt which exist in liquid state at
temperature below 100 8C. Often they are also called room
temperature ionic liquids (RTILs) [60]. For a start, the ionic liquids
possessed greener properties such as very low relative volatility
i.e., close to zero, wide liquids temperature and signicantly less
toxic compared to the organic solvent. More important is their
ability to selectively dissolve various organic, inorganic, and
organometallic materials due to their tuneable polarity which can
be designed for specic purposes. They can also be made miscible
or immiscible with organic solvents and water. Both could be
achieved by simply varying their anioncation combinations.
Because of these, ILs possesses numerous advantages over the
conventional organic solvents, besides being more environmentally compatible [61,62].
In industrial practice, ILs has also been proven through few
chemical manufacturing processes to give signicant advantage
such as in the production of alkoxyphenolphospines which is a
generic precursor for photoinitiator by BASF (BASIL process) where
the reaction time was reduced considerably resulting in tremendous increase in capacity. In addition, the separation of the product
from the catalyst and ionic liquids was also easily attained
compared to the earlier process. Generally for industrial applications, considerable efforts have been made on designing acidic
ionic liquids to replace homogeneous and heterogeneous acids in a
variety of chemical applications [63], such as catalysis in
supporting enzyme catalyzed reactions [64,65] homogeneous
and heterogeneous catalysis [66] purication [67] photo-isomerization [68] and electrochemistry [69]. In view of the environmental friendly nature of ILs, the catalyst developed was also dubbed as
green catalyst [64,65].
Among the major impetus in reaction catalysis developments
involved, is the development of processes in which easy separation
of products and reuse of catalyst is made viable along with high
reactivity and selectivity [70]. In this context, ILs are viewed as
prospective catalysts for biodiesel production where the technical,
economic and environmental aspects were addressed through
their inherent characteristics of being less corrosive, ease of
separation, recyclable, applicable for continuous process and less
production of waste water [71,72]. Applying ILs based catalyst
could reduce the number of reactions and purication steps
required in the biodiesel preparation and separation hence
rendering the process to be more economically competitive. A
typical path way for biodiesel preparation while using ionic liquid
as a catalyst has been depicted in Fig. 2.
The Brnsted acidic ILs possesses several advantageous in
which the uniqueness of combining and designing the solid and
mineral acids to substitute the traditional mineral hazardous
liquid acids, such as H2SO4 [73]. Its applications are widely known

[(Fig._2)TD$IG]

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

O
O
O

R1
R

O
O

OH

R1

O
R2 + 3 ROH

Ionic liquid
R

O
O

R2

OH

O
R3

R3

OH

Fig. 2. Mechanism of biodiesel preparation via ionic liquid catalyzed process.

in the area of esterication with excellent yields and selectivity


[74,75] following its rst synthesis made in 2002 by Cole and coworkers [76]. However, most of the reported works dealt with
aliphatic or arylic esters. Only little information on preparation of
fatty acid alkyl esters using acidic ILs was found in the literature
[51,77].
ILs has also been looked at in addressing the drawbacks
associated with product separation and catalyst recycling [78] in
reactions where for most cases it was impossible to achieve 100%
reaction conversion with 100% selectivity due to the thermodynamic limitations and the competition from parallel reactions [79].
As such, the use of ILs as excellent substitute for traditional
solvents or catalyst in organic synthesis has been growing
considerably along with the prospect of developing green catalysts
[70]. Within the last decade, the potential of ILs applications in the
chemical processes have taken a new uprising turn [79]. Many of
the reported studies have demonstrated the successful application
of ILs as catalysts in hydrogenation, isomerization, CC and CO
cleavage reactions such as catalytic cracking of polyalkenes and C
C coupling reactions such as FriedelCrafts reaction, producing
excellent yields and selectivity [79].
3.1. Ionic liquid coupled with inorganic supporting material as
catalyst for biodiesel reaction
Abreu et al. [80] studied the activity of two multi-phase
systems, for soybean oil alcoholysis based on tin compounds. The
two systems were prepared from the complex Sn(3-hydroxy-2methyl-4-pyrone)2(H2O2) by dissolving it in 1-butyl-3-methylimidazolium hexauorophosphate ([BMIM][PF6]) and supporting it
on acidic resin. On the contrary, their recyclability run results
showed that the multi-phase systems failed to achieve more than
1.0% yield.
Later, DaSilveira et al. [78] immobilized a complex Sn(3hydroxy-2-methyl-4-pyrone)2(H2O)2 in 1-n-butyl-3-methylimidazolium tetrachloro-indate (BMIM+InCl4), and compared their
performance against different catalysts. Their idea in preparing the
biphasic systems in ILs seemed to be based on the expected
advantages for both ILs and heterogeneous catalysis. The various
catalysts tested in their work were lanthanide(III) chloride,
cerium(III) chloride, barium(II), titanium(III), zirconium(IV), copper(II), magnesium(II), zinc(II), cadmium(II), platinum(IV), platinum(II), nickel(II), tin(II) and indium(III) chlorides, aluminum(III)
and germanium(IV) oxides, copper(II) bromide, copper(I) iodide,
Na2O5(H2O), Na2SiF6, and BF3OEt2. However, all of these catalysts
showed very low efciency or almost failed to promote the
reaction, giving only traces of biodiesel. But when the tin complex
was used in replacement of these catalysts, the highest conversion
obtained was reported to be 83% after 4 h of reaction time using
1.0% catalyst. However, the reusability of the catalyst in these two
systems was poor since the yield of biodiesel reduces close to zero

after three cycles. On the other hand, a high biodiesel yield was
also produced from vegetable oils (9398.5%) when an organic
acid/base and a Lewis acid were combined with [Et3NH]Cl-ALCl3
and H2SO4, and are immobilized in [BMIM]NTf2 but there was little
change in the yield shown even after six cycles of re-utilisation
[81]. From the work of Abreu et al. [80] and DaSilveira Neto et al.
[78], it seems that the use of biphasic systems is expensive because
of the number of chemicals used for synthesizing the ILs and for the
preparation of the complex mixtures.
Zhang et al. [82] evaluated the efciency of Brnsted acidic ionic
liquid supported onto Fe-incorporated SBA-15 (Fe-SBA-15) for
esterication of oleic acid with short-chain alcohols. The resultant
catalyst showed high efciency because of the synergistic effect of
Lewis site of Fe and Brnsted acidic sites of ionic liquid. The
esteried product obtained was 87.7% under the optimized
reaction conditions of temperature 363 K, molar ratio of methanol
to oleic acid 6:1, and catalyst amount 5 wt%, and reaction time 3 h.
Yuan et al. [83] evaluated a novel tungstophosphoric anion based
ionic liquid [BsAIm]3[PW12O40], grafted on the surface of silica for
esterication of oleic acid to biodiesel. The yield of 48.49% was
achieved for product.
Zhang et al. [82] evaluated the efciency of Brnsted acidic ionic
liquid supported onto Fe-incorporated SBA-15 (Fe-SBA-15) for
esterication of oleic acid with short-chain alcohols. The resultant
catalyst showed high efciency because of the synergistic effect of
Lewis site of Fe and Brnsted acidic sites of ionic liquid. The
esteried product obtained was 87.7% under the optimized
reaction conditions of temperature 363 K, molar ratio of methanol
to oleic acid 6:1, and catalyst amount 5 wt%, and reaction time 3 h.
Yuan et al. [83] evaluated a novel tungstophosphoric anion based
ionic liquid [BsAIm]3[PW12O40], grafted on the surface of silica for
esterication of oleic acid to biodiesel. The yield of 48.49% was
achieved for product.
Feng et al. [21] evaluated 1-butyl-3-methylimidazolium
tosylate ([BMIm][CH3SO3]) coupled with FeCl3 for transesterication of un-pretreated Jatropha oil containing larger amount of oleic
acid with high-acid value (13.8 mg KOH/g). The yield of biodiesel
was noted to increase from 12% to 93% with raise in temperature
from 120 to 140 8C, when only ionic liquid was used as a catalyst.
Moreover a yield of 99.7% for biodiesel was obtained when FeCl3
was added to the reaction mixture at 120 8C. The high yield at low
temperature was attributed to the metal ions having Lewis acidic
sites, and more of the sites could be provided by trivalent metallic
ions than those of bivalent ones. The mixture of [BMIm][CH3SO3]
and FeCl3 was easily recycled for another turn of esterication.
Young et al. [84] investigated the role of ionic liquid as a cosolvent for acid(HCl)-catalyzed transesterication of biomass
containing intracellular lipids. A yield of 100% was obtained for
all biomass except using Chlorella microalgae. The ionic liquid was
identied to enhance the separation of product from by-products.
It was assumed that ionic liquid containing hydrophobic regions

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

showed some association with the alkyl side chains on the lipid, as
well as provide energy to push the FAME product out of phase with
the co-solvent, and into its own separate and immiscible phase self,
as the ionic liquid anions and cations continued to self-associate.
Zhao et al. [85] synthesized a new type of ether-functionalized
ionic liquids (ILs) carrying anions of acetate or formate; which are
capable of dissolving a variety of substrates and are also lipasecompatible. These ILs are capable of dissolving oils at the reaction
temperature (50 8C); meanwhile, lipases maintained high catalytic
activities in these media even in high concentrations of methanol
(up to 50%, v/v). The study on the transesterication of soybean oil
in IL/methanol mixtures further conrmed the potential of using
oil-dissolving and lipase-stabilizing ILs in the efcient production
of biodiesels.
Tao et al. [86] synthesized Lewis-acidic ionic liquid i.e., choline
chloride  ZnCl2 for transesterication of soybean oil to biodiesel.
The use of this ionic liquid has the advantage of performing the
reaction at mild conditions and also lower cost of synthesizing the
ionic liquid from cheap material.
3.2. Ionic liquid solely as a catalyst for biodiesel reaction
Certain ILs may also be used as catalysts directly for oil
esterication. Wu et al. [51] reported the rst work on direct use
of ILs as transesterication catalyst particularly the acidic ones such
as 1-(4-sulfonic acid) propylpyridiniumhydrogensulfate [HSO3PPyr][HSO4], 1-(4-sulfonic acid) butylpyridiniumhydrogensulfate
[HSO3-BPyr][HSO4], 1-(4-sulfonic acid) propyl-3-methylimidazolium hydrogensulfate [HSO3-PMIm][HSO4], 1-(4-sulfonic acid)
butyl-3-methylimidazolium hydrogensulfate [HSO3-BMIm][HSO4],
and N-(4-sulfonic acid) propyl triethylammoniumhydrogensulfate
[HSO3-PEt3Am][HSO4]. A well rened cottonseed oil was used and
the reactions were catalyzed by various Brnsted acidic ILs
comprising of pyridinium, imidazolium and ammonium as the
cation with butyl and propyl sulfonic acid group as its side chains,
and hydrogen sulfate as the anion (Fig. 3). The ILs 1-(4-sulfonic acid)
butylpyridiniumhydrogensulfate showed the best catalytic performance with a 92% conversion after 5 h at an optimum reaction
conditions of 12:1 oil to ILs molar ratio, catalyst concentration of
5.7 wt%, and temperature at 170 8C.
[BSPy]CF3SO3 showed consistent activity after seven successive
cycles in the transesterication reaction of Jatropha oil [87]. More
recently, Han et al. [77] prepared biodiesel from waste oil using
acidic [HSO3-BPyr][HSO4] as a catalyst. The yield was found to be
93.6% after 4 h of reaction time using 6:1 oil to ILs molar ratio, 6%
catalyst concentration and temperature at 170 8C.

[(Fig._3)TD$IG]

Fig. 3. Effect of IL type on the FFA conversion to methyl esters (methanol/CPO molar
ratio 21:1, temperature 170 8C, and rate of stirring 600 rpm for 4 h) [88].

Anions play a pivotal role in the acidic nature of ILs [82], and the
substituents on cations may have an important inuence on the
accessibility of the acid active sites. Ghiaci et al. [87] studied
the effect of benzyl side chain (attached to cation) on acidity of
ionic liquid for transesterication of canola oil with methanol. The
results show that the 4B ionic liquid has the highest catalytic
activity and best recyclability under the optimized reaction
conditions.
Elsheikh et al. [88] evaluated three imidazolium-based ionic
liquids containing different size of alkyl side chains on cation and
xed acidic anion HSO4. These three ionic liquids were BMIMHSO4,
BIMHSO4 and MIMHSO4. The BMIMHSO4 ionic liquid due to longer
alkyl side chain was found more efcient than the other two ionic
liquids (Fig. 3). The high catalytic activity of BMIMHSO4 was
correlated to its acidity which was attributed to the steric
hindrance of alkyl side chains on cation moiety. The optimum
conditions for highest yield was 4.5 wt.% loading of BIMHSO4,
methanol/CPO molar ratio of 12:1, a temperature of 160 8C, and
agitation speed of 600 rpm.
Similarly Liu et al. [89] employed 1-(3-sulfonic acid) propyl-3methylimidazolehydrosulfate[HO3S-pmim]HSO4 for transesterication of waste oil. The yield of biodiesel was more than 96% for
the reaction under the optimum conditions: oil/methanol ratio
1:12, waste oil 15.0 g, ionic liquid 2.0 g, reaction temperature
120 8C and pretreatment time 8 h. The ionic liquid was recycled
ve times with more than 93% of biodiesel yield (Fig. 4).
Elsheikh et al. [90] also studied pyrazoliums based ionic liquids
i.e., sulfoalkyl-pyrazoliumhydrogensulfate and alkylsulfo-alakylpyrazoliumhydrogensulfate for esterication of bitter apple oil.
The results showed that 2-(4-sulfobutyl) pyrazoliumhydrogensulfate (SBPHSO4) ionic liquid due to strong acidity, was good in its
catalytic activity. A yield of 89.5% was obtained when reaction was
carried under the condition of 5.2 wt% of SBPHSO4, molar ratio of
methanol to BAO of 15:1, 170 8C, and 800 rpm for 6 h.
Liu et al. [91] investigate the ionic liquid [HO3S(CH2)3
NEt3]ClFeCl3 containing both Brnsted and Lewis acidic sites for
synthesis of biodiesel from waste cooking oil. Due to dual
functionality in the ionic liquid the yield of biodiesel was obtained
even more than 95% at 120 8C for 4 h. Man et al. [92] used
triethylammoniumhydrogensulfate (Et3NHSO4) as pretreatment
solvent followed by KOH for transesterication of crude palm oil.
The biodiesel in the yield of 82.1% was obtained when only 5.2 wt.%
(with respect to reaction mixture) of Et3NHSO4 was used at 170 8C
for 3 h. The end product obtained after KOH treatment was 96.9%.
Fang et al. [93] compared a basic ionic liquid with conventional
catalysts for transesterication of castor oil with methanol and
high
yield of biodiesel i.e., 96% was achieved under optimum
[(Fig._4)TD$IG]

Fig. 4. Reusability of IL [89].

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

reaction conditions. Zhang et al. [94] investigated the N-methyl-2pyrrolidonium methyl sulfonate ([NMP][CH3SO3]) as a catalyst
without addition of other organic acids for transesterication
reaction of fatty acids. The yield of fatty acid alkyl esters of 93.6%
and 95.3% was obtained at 70 8C for 8 h the ionic liquids was
effective in catalysis even recycled after eight times.
Liang et al. [30] synthesized acidic-based solid ionic liquid
polymer by coupling acidic ionic liquid oligomers with divinylbenzene. The synthesized ionic liquid was used as a catalyst for
transesterication of triglycerides, a total yield of >99% of
biodiesel was attained in one spot reaction under mild conditions.
The high acidity of synthesized ionic liquid was attributed to its
high hydrophobic surface area, high acidity, and high stability.
Guo et al. [23] investigated the effect of ultrasound irradiation
on transesterication of soybean oil with methanol in the presence
of a Brnsted acidic ionic liquid. Under predetermined conditions
of methanol/oil molar ratio 9:1, 1.0 wt.% catalyst in oil, ultrasound
power of 200 W, and temperature of 60 8C, the yield was about
93.2% within the reaction time of 60 min. Cui et al. [95] studied the
kinetic behavior of ionic liquid catalyzed transesterication of
methyl acetate with n-butanol. Two different kinetic models, the
ideal homogeneous (IH) model and the non-ideal homogeneous
(NIH) model, were used to correlate the kinetic data. The ionic
liquid catalysis was compared with conventional catalysts such as
ion-exchange resin catalysts sulfuric acid and Amberlyst 15. The
results showed that ionic liquid catalysis was effective than the
two conventional catalysts.
Liang et al. [27] reported dication ionic liquids of imidazolium
based with x Brnsted basic anion of OH for transesterication
of cottonseed oil. The best result of biodiesel preparation with the
yield of 98.5% was obtained for bis-(3-methyl-1-imidazolium)ethylene dihydroxide at 55 8C, 0.4% catalyst and 12:1 methanol to
cottonseed oil molar ratio after 4 h of reaction time. He et al. [96]
synthesis and characterize a long-chain Brnsted acid ionic liquid
(IL), 3-(N,N-dimethyldodecylammonium)propanesulfonic acid ptoluenesulfonate ([DDPA][Tos]) for synthesis of biodiesel from free
fatty acids (FFAs). The product of biodiesel was obtained with good
yield of 92.596.5% under optimum condition of: molar ratio of
alcohols to FFA at 1.5:1, mole fraction of ionic liquid at 10%, 60 8C,
and 3 h. The long-chain Brnsted acidic ionic liquid was good in
catalysis and could be recycled up to nine times.
3.3. Ionic liquid as deep eutectic solvents (DES) for biodiesel reaction
Recently, a special class of ILs known as the deep eutectic
solvents (DES), was introduced and they are believed to be a new
generation of ionic liquids. The DES has received increasing
interest due to their potential to be even more environmentally
benign compared to the earlier traditional ILs besides also
possessing equally good if not better solvation properties.
Zhao and Baker [97] critically addressed the possibility of
combining the conventional ILs with deep eutectic solvents for
biodiesel production. The DES comprise of mixtures of organic
halide salts, such as choline chloride with an organic compound
that is a hydrogen bond donor (HBD) capable of forming a
hydrogen bond with the halide ion, such as amides, amines,
alcohols, carboxylic acids and many more [98]. The liquid state of
the DES is produced through freezing point depression, whereby
hydrogen-bonding interactions between an anion and an HBD are
more energetically favored relative to the lattice energies of the
pure constituents [99]. In another study, Huang et al. [100]
identied a straightforward and energy efcient method to
activate commercial CaO for biodiesel production without any
pretreatment by the addition of a novel DES which could easily
remove the inactive layers of calcium carbonate and calcium
hydroxide on the surface of the commercial CaO during the

reaction to achieve a high FAME yield, while the reaction resulted


in an extremely low yield without the addition of DES. Among the
main advantages of the DES over the traditional ILs are that they
are easier to be prepared in pure state and they are not reactive to
water. Studies on its toxicology reveal that quite a signicant
number of them are biodegradable [101].
Hayyan et al. [102] used a phosphonium-based deep eutectic
solvent (P-DES) followed by alkali treatment for esterication of
low grade crude palm oil (LGCPO). The P-DES was added in the
range from 0.25 to 3.5% (wt/wt) of reaction mixture and high
catalytic activity in the pre-treatment of LGCPO was noted. Hayyan
et al. [103] also evaluated the ammonium-based deep eutectic
solvent which consisted of hydrogen bond donor (i.e., ptoluenesulfonic acid monohydrate) (PTSA) and salt (i.e., N,Ndiethylenethanol ammonium chloride) as a novel recyclable
catalyst (DEAC-DES) for transesterifcation of LGCPO. The free fatty
acid content of LGCPO was reduced from 9.5% to 1% using optimum
conditions, could be efciently converted to biodiesel after alkali
treatment.
ILs provide an ideal medium for the separation of the biodiesel
and for the removal of the glycerol by-product. However, there is a
lack of long-term and/or continuous biodiesel production systems
based on ILs, and the accumulation of glycerol and the recovery of
the ILs are likely to pose future challenges. In addition, the
environmental fate and any potential toxicity issues for most ILs
are still largely unknown and up to the present, only a few
preliminary reports on the toxicological properties of them were
available.
4. Role of ionic liquid as a solvent for biocatalyzed
transesterication process
The enzymatic transesterication method have many advantages over chemical methods such as mild reaction conditions
leading to low energy consumption, exibility in choosing
different enzymes for various types of substrates, reusability of
enzymes, reduced water requirement in substrates hence lowering
the waste treatment, etc. [54]. Unfortunately, the current lipasecatalyzed method exhibits several downsides which prevent it
from being commercialized. Among others, these include the high
enzymes cost and lipase inactivation by acyl acceptors such as
methanol and by impurities in crude and waste oils. In addition,
due to the poor miscibility between oils/fats with methanol, many
of the enzymatic transesterication reactions have to be conducted
in heterogeneous systems involving complicated liquidliquid
interface [54].
On the issue of methanol inhibition of lipases, several
approaches have been introduced such as stepwise addition of
methanol during reaction [104], use of other acyl acceptors such as
methyl and ethyl acetate [105], enzyme immobilization, use of
other organic solvents such as t-butanol, hexane, n-heptane, and
1,4-dioxane, use of fatty acid-containing feedstock [105], and
genetic modication of lipases for higher methanol tolerance [54].
Nevertheless, some of the alternative solvents are found to be
toxic to the lipase enzyme resulting in poor enzyme activity and
stability [106]. Besides, it was also observed that the lipase can be
deactivated by glycerol which is a byproduct from the transesterication reaction through its adsorption onto the enzyme surface
[107]. Thus, in order to produce biodiesel using a truly green
process, a combination of catalysts and solvents that are both
environmentally friendly are needed.
Due to their unique and greener properties, ILs has been
identied as a promising option to meet the requirement. The
enzymatic transesterication of vegetable oils in ionic liquids has
been demonstrated by several groups in producing biodiesel
[107,108,39,109]. Placing the enzymes in the imidazolium type ILs

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

has been found to generally produce more active and stable


catalysts as the ILs were able to protect the lipase from
deactivation induced by methanol [110]. A schematic representation of enzymatic biodiesel process in ILs is shown in Fig. 5.
Typically, the ILs containing shorter chain 1,3-dialkylimidazolium cation (e.g., [BMIM]PF6 or [BMIM]NTf2) were used and the
reaction was performed in a biphasic system which requires a
certain amount of water. The ILs [BMIM]NTf2, when used in
combination with lipase for biodiesel production, was reported to
produce a biodiesel yield of 96.3% from soybean oil [111].
In another study conducted by Yang et al. [108] involving the
addition of a salt hydrate directly to the non-aqueous system, was
found to be unsuitable to meet the water requirement of the
lipase-catalyzed reaction in [BMIM]PF6. The salt hydrate may
affect both the water buffering and the specic ion effect on the
enzyme activities but the latter was found to be more dominant in
the IL system [92]. When the water is added to the originally
biphasic system, it causes a formation of a third phase. Normally,
after transesterication, the water phase will help in capturing the
by-products and the acyl acceptor, thereby facilitating the recovery
and reuse of the ILs. However, it was shown in a study that the
lipase activity demonstrated a decline in its efciency after the
sixth cycle of reused [112].
Contrary to the use of shorter alkyl chain cation, the
Imidazolium ILs containing longer alkyl chains (e.g., [C16MIM]NTf2
and [C18MIM]NTf2) have been used in a homogeneous one-phase
system for lipase-catalyzed biodiesel production. The benet of
this is that it protects the enzyme and avoids direct interactions
with methanol thus allowing for the stable reuse of lipase in the ILs
[11,110]. These long chain and lipophilic ILs create a nonaqueous
system for oil transesterication, and a triphasic system at the end
[(Fig._5)TD$IG]of the reaction, facilitates biodiesel extraction [11,110]. Similar to

the inuence of ILs on cellulase activity, increasing the viscosity of


the ILs with [CnMIM] cations produced a negative effect on the
lipase-catalyzed transesterication, in an anion order that was
consistent with increasing viscosity: [PF6] > [BF4] > [NTf2] [110].
Ha et al. [113] has screened 23 ILs for methanolysis of soybean
oil catalyzed by immobilized Candida antarctica lipase (Novozym1
435), and identied the hydrophilic IL (i.e., [EMIM][OTf]) as the
best solvent for achieving the highest yield (80%) of fatty acid
methyl esters based on 12 h reaction time. On the other hand,
Sunitha et al. [114] obtained 9899% yields of fatty acid methyl
esters within 10 h of methanolysis of sunower oil in hydrophobic
[BMIM][PF6] and [EMIM][PF6] when catalyzed by Novozym 435
(Table 1). Zhaoa et al. [115] introduced choline-based deep eutectic
solvents (such as choline chloride/glycerol at 1:2 molar ratio) as
inexpensive, non-toxic, biodegradable and lipase-compatible
solvents for the enzymatic preparation of biodiesel from soybean
oil. Through the evaluation of different eutectic solvents and
different lipases, as well as the study of reaction parameters (i.e.,
methanol concentration, Novozym 435 loading and reaction time),
they achieved up to 88% triglyceride conversions in 24 h.
Gamba et al. [111] also utilized several ILs (i.e., [BMIM] [Tf2N],
[BMIM][BF4], and [BMIM][PF6]) as solvents in the enzymatic
transesterication of soybean oil, achieving over 90% biodiesel
yield in 48 h. However, hydrophobic ILs usually does not dissolve
triglycerides and the lipase, resulting in a multi-phase reaction
[111,113]. On the other hand, hydrophilic ILs often causes enzyme
dissolution and denaturation [15,116].
Rhizopusoryzae and Aspergillusoryzae were used as a wholecell biocatalyst for biodiesel production in ionic liquids due to their
lipase-production ability. In this system, the whole-cells can be
used as a lipase-immobilized matrix, and ionic liquids play a role in
avoiding deactivation of the biocatalysts. The use of ILs, which is

Fig. 5. Enzymatic biodiesel process in ionic liquids [17].

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

Table 1
Effect of concentrations and types of ionic liquids on Candida antarctica-catalyzed methanolysis of sunower oila [114].
Ionic liquidb

Oil:ionic liquid (w/w)

Relative product composition (% w/w)c


FAME

TG

DG

MG

BMIm PF6

1:2
1:1
4:3
2:1
4:1
1:2
1:1
4:3
2:1
4:1
1:2
1:2

97  0.8
98  0.6
95  1
38  1.5

98  0.8
98  1.6
96  1.4
46  3

10  2.4

1  0.8

31
50  2
98  0.8

21
40  3
99  1
80  2.5
94  1

2  0.3
1  0.2

8  0.81
1  0.8
1  0.8
1  1.6
1  1.4
61

32
51

1  0.2

1  0.2

EMIm PF6

HMIm BF4
BMIm BF4

a
Reaction conditions: sunower oil (1 g, average equivalence 1.3 mM), methanol (0.4 ml, 10.4 mM), Candida antarctica (10%, 0.1 g) at 60 8C for 4 h. Values are reported as the
mean of three independent reactions  SD.
b
[BMIm][BF4], 1-butyl-3-methyl imidazolium tetrauoroborate; [BMIm][PF6], 1-butyl-3-methyl imidazolium hexauorophosphate; [EMIm][PF6], 1-ethyl-3-methyl
imidazolium hexauorophosphate; [HMIm][BF4], 3-methyl imidazolium tetrauoroborate.
c
FAME, fatty acid methyl ester; TG, triacylglycerol; DG, diacylglycerol; MG, monoacylglycerol.

totally composed of salt, make the recovery of methyl ester easy as


it does not dissolve the methyl esters and the unreacted oil thus
forming a biphasic system. Moreover, ionic liquids can also be
expected to act as an extracting phase for the by-product glycerol,
which often has a negative effect on lipase activity in biodiesel
production [117,118].
Lipase PS i.e., Burkholderiacepacia lipase (BCL) has been widely
applied for biotransformation reactions both in aqueous and nonaqueous phases [119,120]. BCL has been immobilized on NKA resin
to produce a high performance biocatalyst which could be used for
synthesizing biodiesel in isooctane and solvent free systems
[121,122]. Similarly when BCL is coated with RTILs, it was noted
that the enantioselectivity for various substrates was tremendously improved [123].
Lozano et al. [124] investigated the ILs N-octadecyl-N0 ,N00 ,N000 trimethylammoniumbis(triuoromethylsulfonyl)imide
([C18tma][NTf2]) coupled with lipase for methanolysis of triolein to
obtain biodiesel. The reaction was divided into liquid and solid
phases to evaluate the yield and recycling of reaction components.
The reaction carried out in liquid phases has better results and the
yield obtained was 100% at temperature of 60 8C with pretreatment time of 8 h. Enhanced enzyme stability of up to 1370
days half-life time was observed for the ionic liquid at 60 8C. In case
of solid phases the reaction mixture was completely separated into
solid IL, glycerol and pure biodiesel via centrifugation at controlled
temperature (Fig. 6).
Diego et al. [125] used hydrophobic ionic liquids (containing
Cation; [C10-C18MIM], Anion; [BF4], [PF6] or [NTf2]) as reaction
media for biocatalysis resulting in the biodiesel production. Two
types of lipases i.e., (C. antarctica lipase B and Pseudomonas
uorescens lipase AK) were analyzed. The highest activity was
noted when coupling the Novozym 435 (immobilized C. antarctica
lipase with [C16MIM] [NTf2] ILs). The activity in the presence of the
ILs was three times more as compared to the solvent-free system
and the biodiesel yield after 3 h of reaction time was 90.29% and
27.3% respectively for the case with and without the ILs.
Pan et al. [126] evaluated the esterication activity of Novozym
435 and Lypozyme RM IM in ten different ionic liquids for the
synthesis of tricaprylin. Novozym 435 was observed to have higher
catalytic activity compared to Lypozyme RM IM for all the studied
ILs. Among the ILs used, those containing Tf2N and PF6 anions
shown higher catalytic activity than BF4 anion. Subsequently,
they also possessed higher activity compared to the system using
hexane as the solvent. The higher catalytic activity of Novozym 435
observed in [C4MIM] Tf2N and [C4MIM] PF6 was argued to be due to

the reduction in a-helix-secondary structure of the lipase which


was supported through analysis conducted on FTIR. A higher yield
of 92.4% for tricaprylin was obtained under the optimum reaction
condition i.e., the lipase amount of 6.1% substrate mass at a caprylic
acid/glycerol molar ratio of 4.5:1 and temperature of 66.7 8C.
Lai et al. [25] extract lipids from microalgae Botryococcus braunii
strains, Chlorella vulgaris (CV), and Chlorella pyrenoidosa (CP) and
were esteried by two immobilized lipases, Penicillium expansum
lipase (PEL) and C. antarctica lipase B (Novozym 435) in solvent
systems of 1-butyl-3-methylimidazolium hexauorophosphate,
[BMIm][PF6] ionic liquid and tert-butanol. In the predetermined
conditions, both enzymes induced signicantly higher yields in the
IL (90.7% and 86.2%) relative to that obtained in tert-butanol (48.6%
and 44.4%). In comparison, PEL results slightly in higher yield than
Novozym 435.
The range of examples demonstrated shown the fact that ILs has
the potential of a supporting and good reaction medium for
biodiesel production. However, it has to be highlighted here that
the Novozym 435 has mainly always been used as the catalyst for
all these applications.
5. Conclusion

[(Fig._6)TD$IG]

Biodiesel production from non-edible oil is one of the potential


alternative energy sources which would reduce the world
dependency on petroleum based diesel for generation of energy.
The ionic liquids especially those having acidic nature show good
catalytic activity for biodiesel production. The ionic liquid
pretreatment has the advantage over other techniques as the
resulted biodiesel and glycerol could be easily separated. Also, it

Fig. 6. Fractionated reaction mixture of solid IL, glycerol and pure biodiesel [124].

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

has been shown that the ionic liquids could be recycled many
times without decrease in efciency. Ionic liquids have also been
tested as solvent for catalytic reactions of biodiesel production.
Both inorganic and organic catalysts were added to enhance the
catalytic activity of the ionic liquid. Generally, the ILs especially the
hydrophobic ones were found to be good solvent for biocatalysis as
it will not denature the enzyme.
Acknowledgement
The authors wish to thank Higher Education Commission,
Government of Pakistan for providing nancial support for the
current study under the National Research Program for Universities (NRPU).
References
[1] Y. Zhang, M.A. Dube, D.D. McLean, M. Kates, Bioresource Technology 89 (2003) 1.
[2] A.K. Endalew, K. Kiros, R. Zanzi, Biomass and Bioenergy 35 (2011) 3787.
[3] W.J. Ting, C.M. Huang, N. Giridhar, W.T. Wu, Journal of the Chinese Institute of
Chemical Engineers 39 (2008) 203.
[4] M. Verziu, B. Cojocaru, J. Hu, R. Richards, C. Ciuculescu, P. Filip, V.I. Parvulescu,
Green Chemistry 10 (2008) 373.
[5] M.M. Soumanou, U.T. Bornscheuer, European Journal of Lipid Science and
Technology 105 (2003) 656.
[6] Z. Masoud, A.W.D. Wan Mohd, K. Mohamed, Fuel Processing Technology 90
(2009) 770.
[7] N. Jain, A. Kumar, S. Chauhan, S.M.S. Chauhan, Tetrahedron 61 (2005) 1015.
[8] F. van Rantwijk, R.A. Sheldon, Chemical Reviews 107 (2007) 2757.
[9] T. De Diego, P. Lozano, S. Gmouh, M. Vaultier, J.L. Iborra, Biotechnology and
Bioengineering 88 (2004) 916.
[10] J.L. Kaar, A.M. Jesionowski, J.A. Berberich, R. Moulton, A.J. Russell, Journal of the
American Chemical Society 125 (2003) 4125.
[11] P. Lozano, E. Garca-Verdugo, R. Piamtongkam, N. Karbass, T. De Diego, M.I.
Burguete, S.V. Luis, J.L. Iborra, Advanced Synthesis & Catalysis 349 (2007) 1077.
[12] K.-W. Kim, B. Song, M.-Y. Choi, M.-J. Kim, Organic Letters 3 (2001) 1507.
[13] H.-P. Zhu, F. Yang, J. Tang, M.-Y. He, Green Chemistry 5 (2003) 38.
[14] H. Pfruender, M. Amidjojo, U. Kragl, D. Weuster-Botz, Angewandte Chemie
International Edition 43 (2004) 4529.
[15] R. Madeira Lau, M.J. Sorgedrager, G. Carrea, F. van Rantwijk, F. Secundo, R.A.
Sheldon, Green Chemistry 6 (2004) 483.
[16] http://www.palmoilhq.com.
[17] T.D. Diego, A. Manjon, P. Lozano, J.L. Iborra, Bioresource Technology 102 (2011)
6336.
[18] Y.A. Elsheikh, Industrial Crops and Products 49 (2013) 822.
[19] M. Fan, J. Huang, J. Yang, P. Zhang, Applied Energy 108 (2013) 333.
[20] M.M. Fan, J.J. Zhou, Q.J. Han, P.B. Zhang, Chinese Chemical Letters 23 (2012) 1107.
[21] F. Guo, Z. Fang, X.-F. Tian, Y.-D. Long, L.-Q. Jiang, Bioresource Technology 102
(2011) 6469.
[22] F. Guo, Z. Fang, X.-F. Tian, Y.-D. Long, L.-Q. Jiang, Bioresource Technology 140
(2013) 447.
[23] W. Guo, H. Li, G. Ji, G. Zhang, Bioresource Technology 125 (2012) 332.
[24] M. Hayyan, F.S. Mjalli, M.A. Hashim, I.M. AlNashef, Fuel Processing Technology
91 (2010) 116.
[25] J.-Q. Lai, Z.-L. Hu, P.-W. Wang, Z. Yang, Fuel 95 (2012) 329.
[26] D.Y.C. Leung, Y. Guo, Fuel Processing Technology 87 (2006) 883.
[27] J.-h. Liang, X.-q. Ren, J.-t. Wang, m. Jinag, Z.-j. Li, Journal of Fuel Chemistry and
Technology 38 (2010) 275.
[28] M. Di Serio, R. Tesser, M. Dimiccoli, F. Cammarota, M. Nastasi, E. Santacesaria,
Journal of Molecular Catalysis A: Chemical 239 (2005) 111.
[29] X. Liang, Applied Catalysis A: General 455 (2013) 206.
[30] X. Liang, H. Xiao, C. Qi, Fuel Processing Technology 110 (2013) 109.
[31] Y. Liu, D. Chen, Y. Yan, C. Peng, L. Xu, Bioresource Technology 102 (2011) 10414.
[32] T. Long, Y. Deng, S. Gan, J. Chen, Chinese Journal of Chemical Engineering 18
(2010) 322.
[33] M.J. Haas, A.J. McAloon, W.C. Yee, T.A. Foglia, Bioresource Technology 97 (2006)
671.
[34] J.V. Gerpen, Fuel Processing Technology 86 (2005) 1097.
[35] E. Crabbe, C. Nolasco-Hipolito, G. Kobayashi, K. Sonomoto, A. Ishizaki, Process
Biochemistry 37 (2001) 65.
[36] M.A. Kalam, H.H. Masjuki, Biomass and Bioenergy 23 (2002) 471.
[37] B.S. Souza, D.M.M. Pinho, E.C. Leopoldino, P.A.Z. Suarez, F. Nome, Applied
Catalysis A: General 433434 (2012) 109.
[38] Q. Wu, H. Wan, H. Li, H. Song, T. Chu, Catalysis Today 200 (2013) 74.
[39] K.-P. Zhang, J.-Q. Lai, Z.-L. Huang, Z. Yang, Bioresource Technology 102 (2011)
2767.
[40] E. Lotero, Y. Liu, D.E. Lopez, K. Suwannakarn, D.A. Bruce, J.G. Goodwin, Industrial
& Engineering Chemistry Research 44 (2005) 5353.
[41] M.C. Math, S.P. Kumar, S.V. Chetty, Energy for Sustainable Development 14
(2010) 339.
[42] M.K. Lam, K.T. Lee, A.R. Mohamed, Biotechnology Advances 28 (2010) 500.

[43] U. Schuchardt, R. Sercheli, R.M. Vargas, Journal of the Brazilian Chemical Society
9 (1998) 199.
[44] M. Formo, Journal of the American Oil Chemists Society 31 (1954) 548.
[45] B. Freedman, E.H. Pryde, T.L. Mounts, Journal of the American Oil Chemists
Society 61 (1984) 1638.
[46] G. Vicente, M. Martnez, J. Aracil, Bioresource Technology 92 (2004) 297.
[47] H. Fukuda, A. Kondo, H. Noda, Journal of Bioscience and Bioengineering 92 (2001)
405.
[48] Y. Liu, E. Lotero, J.G. Goodwin Jr., C. Lu, Journal of Catalysis 246 (2007) 428.
[49] A.A.M. Lapis, L.F. dejOliveira, B.A.D. Neto, J. Dupont, ChemSusChem 1 (2008) 759.
[50] H. He, T. Wang, S. Zhu, Fuel 86 (2007) 442.
[51] Q. Wu, H. Chen, M. Han, D. Wang, J. Wang, Industrial & Engineering Chemistry
Research 46 (2007) 7955.
[52] X. Liang, G. Gong, H. Wu, J. Yang, Fuel 88 (2009) 613.
[53] H. Fukuda, S. Hama, S. Tamalampudi, H. Noda, Trends in Biotechnology 26 (2008)
668.
[54] C.C. Akoh, S.W. Chang, G.-C. Lee, J.-F. Shaw, Journal of Agricultural and Food
Chemistry 55 (2007) 8995.
[55] M. Berrios, R.L. Skelton, Chemical Engineering Journal 144 (2008) 459.
[56] M. Canakci, J.V. Gerpen, Transactions of the ASAE 44 (2001) 1429.
[57] S. Cooke, S. Abrams, B. Bertram, Purication of biodiesel with adsorbent materials, US, 2003.
[58] C.W. Chiu, M.A. Dasari, W.R. Sutterlin, G.J. Suppes, Industrial and Engineering
Chemistry Research 45 (2006) 791.
[59] J.C. Yori, S.A. DIppolito, C.L. Pieck, C.R. Vera, Energy and Fuels 21 (2007) 347.
[60] T. Welton, Chemical Reviews 99 (1999) 2071.
[61] Shariati, K. Gutkowski, C.J. Peters, AIChE Journal 51 (2005) 1532.
[62] J.F. Brennecke, E.J. Maginn, AIChE Journal 47 (2001) 2384.
[63] H. Xing, T. Wang, Z. Zhou, Y. Dai, Molecular Catalysis A: Chemical 264 (2007) 53.
[64] J. Shen, H. Wang, H. Liu, Y. Sun, Z. Liu, Journal of Molecular Catalysis A: Chemical
280 (2008) 24.
[65] J. Gui, X. Cong, D. Liu, X. Zhang, Z. Hu, Z. Sun, Catalysis Communications 5 (2004)
473.
[66] J. Gorman, Science News 160 (2001) 156.
[67] S.G. Cull, J.D. Holbrey, V. Vargas-Mora, K.R. Seddon, G.J. Lye, Biotechnology and
Bioengineering 69 (2000) 227.
[68] X. Li, D. Zhao, Z. Fei, L. Wang, Science in China Series B 49 (2006) 385.
[69] T. Tsuda, C.L. Hussey, Interface-Electrochemical Society 16 (2007) 42.
[70] M. Picquet, D. Poinsot, S. Stutzmann, I. Tkatchenko, I. Tommasi, P. Wasserscheid,
J. Zimmermann, Topics in Catalysis 29 (2004) 139.
[71] M.J. Earle, K.R. Seddon, Pure and Applied Chemistry 72 (2000) 1391.
[72] C. Xie, H. Li, L. Li, S. Yu, F. Liu, Journal of Hazardous Materials 151 (2008) 847.
[73] D. Fang, K. Gong, Q. Shi, Z. Liu, Catalysis Communications 8 (2007) 1463.
[74] D.C. Forbes, K.J. Weaver, Journal of Molecular Catalysis A: Chemical 214 (2004)
129.
[75] Z. Zhang, W. Wu, B. Han, T. Jiang, B. Wang, Z. Liu, The Journal of Physical
Chemistry B 109 (2005) 16176.
[76] A.C. Cole, J.L. Jensen, I. Ntai, K.L.T. Tran, K.J. Weaver, D.C. Forbes, J.H. Davis, Journal
of the American Chemical Society 124 (2002) 5962.
[77] M. Han, W. Yi, Q. Wu, Y. Liu, Y. Hong, D. Wang, Bioresource Technology 100
(2009) 2308.
[78] B.A. DaSilveira Neto, M.B. Alves, A.A. Lapis, F.M. Nachtigall, M.N. Eberlin, J.
Dupont, P.A. Suarez, Journal of Catalysis 249 (2007) 154.
[79] D. Zhao, M. Wu, Y. Kou, E. Min, Catalyst Today 74 (2002) 157.
[80] F.R. Abreu, M.B. Alves, C.S. Macedo, L.F. Zara, P.A.Z. Suarez, Molecular Catalysis A:
Chemical 227 (2005) 263.
[81] Teresa De Diego, Arturo Manjon, Pedro Lozano, J.L. Iborra, Bioresource Technology 102 (2011) 6336.
[82] L. Zhang, Y. Cui, C. Zhang, L. Wang, H. Wan, G. Guan, Industrial & Engineering
Chemistry Research 51 (2012) 16590.
[83] L. Yuan, Z. Bin, L. Hansheng, Industrial Catalysis 19 (2011) 27.
[84] G. Young, F. Nippen, S. Titterbrandt, M.J. Cooney, Biofuels 2 (2011) 261.
[85] H. Zhao, Z. Song, O. Olubajo, J.V. Cowins, Applied Biochemistry and Biotechnology 162 (2010) 13.
[86] L. Tao, Y. Deng, S. Gan, C. Ji, Chinese Journal of Chemical Engineering 18 (2010)
322.
[87] M. Ghiaci, B. Aghabarari, S. Habibollahi, A. Gil, Bioresource Technology 102
(2011) 1200.
[88] Y.A. Elsheikh, Z. Man, M.A. Bustam, S. Yusup, C.D. Wilfred, Energy Conversion and
Management 52 (2011) 804.
[89] S. Liu, Z. Wang, S. Yu, C. Xie, Bulletin of the Chemical Society of Ethiopia 27 (2013)
289.
[90] Y.A. Elsheikh, Process Safety and Environmental Protection (2013).
[91] S. Liu, Z. Wang, K. Li, L. Li, S. Yu, F. Liu, Z. Song, Renewable Sustainable Energy 5
(2013) 1.
[92] Z. Man, Y.A. Elsheikh, M.A. Bustam, S. Yusup, M.I.A. Mutalib, N. Muhammad,
Industrial Crops and Products 41 (2013) 144.
[93] D. Fang, C. Jiang, J. Yang, Energy Technology 1 (2013) 135.
[94] L. Zhang, M. Xian, Y. He, L. Li, J. Yang, S. Yu, X. Xu, Bioresource Technology 100
(2009) 4368.
[95] X. Cui, J. Cai, Y. Zhang, R. Li, T. Feng, Industrial & Engineering Chemistry Research
50 (2011) 11521.
[96] L. He, S. Qin, T. Chang, Y. Sun, X. Gao, Catalysis Science & Technology 3 (2013)
1102.
[97] H. Zhao, G.A. Baker, Journal of Chemical Technology & Biotechnology 88
(2013) 3.

10

N. Muhammad et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 110

[98] L. Liu, Y. Kong, J.P.L.H. Xu, Z.L.J.X. Dong, Microporous and Mesoporous Materials
115 (2008) 624.
[99] C.A. Nkuku, R.J. LeSuer, Journal of Physical Chemistry 111 (2007) 13271.
[100] W. Huang, S. Tang, H. Zhao, S. Tian, Industrial & Engineering Chemistry Research
52 (2013) 11943.
[101] A.P. Abbott, D. Boothby, G. Capper, D.L. Davies, R.K. Rasheed, Journal of the
American Chemical Society 126 (2004) 9142.
[102] A. Hayyan, M. Ali Hashim, F.S. Mjalli, M. Hayyan, I.M. AlNashef, Chemical
Engineering Science 92 (2013) 81.
[103] A. Hayyan, M.A. Hashim, M. Hayyan, F.S. Mjalli, I.M. AlNashef, Industrial Crops
and Products 46 (2013) 392.
[104] Y. Shimada, Y. Watanabe, A.T.Y. Sugihara, Journal of Molecular Catalysis B:
Enzymatic 17 (2002) 133.
[105] W. Du, Y. Xu, D. Liu, J. Zeng, Journal of Molecular Catalysis B: Enzymatic 30
(2004) 125.
[106] X.Q. Chen, C.J. Liu, J. Wang, Y. Li, Chinese Journal of Organic Chemistry 29 (2009)
128.
[107] A.P. de los Ros, F.J.H. Fernandez, D. Gomez, M. Rubio, G. Vllora, Process
Biochemistry 46 (2011) 1475.
[108] Z. Yang, K.-P. Zhang, Y. Huang, Z. Wang, Journal of Molecular Catalysis B:
Enzymatic 63 (2010) 23.
[109] N.-W. Li, M.-H. Zong, H. Wu, Process Biochemistry 44 (2009) 685.
[110] T. De Diego, P. Lozano, S. Gmouh, M.I.,J.L. Vaultier, Biomacromolecules 6 (2005)
1457.
[111] M. Gamba, AAM.D.J. Lapis, Advanced Synthesis and Catalysis 350 (2008) 160.

[112] N.I. Ruzich, A.S. Bassi, The Canadian Journal of Chemical Engineering 88 (2010)
277.
[113] H.S.H.M.N. Lan, S.H. Lee, S.M. Hwang, Y.-M. Koo, Enzyme and Microbial Technology 41 (2007) 480.
[114] S. Sunitha, S. Kanjilal, P.S. Reddy, R.B.N. Prasad, Biotechnology Letters 29 (2007)
1881.
[115] H. Zhao, C. Zhang, T.D. Crittle, Journal of Molecular Catalysis B: Enzymatic 8586
(2013) 243.
[116] H. Zhao, G.A. Baker, S. Holmes, Organic & Biomolecular Chemistry 9 (2011) 1908.
[117] A.P. Abbott, P.M. Cullis, M.J. Gibson, R.C. Harris, E. Raven, Green Chemistry 9
(2007) 868.
[118] V. Dossat, D. Combes, A. Marty, Enzyme and Microbial Technology 25 (1999) 194.
[119] L.H. Andrade, L.P. Rebelo, C.G.C.M. Netto, H.E. Toma, Journal of Molecular
Catalysis B: Enzymatic 66 (2010) 55.
[120] Z.G. Chen, R.X. Tan, M. Huang, Process Biochemistry 45 (2010) 415.
[121] T. Liu, Y. Liu, X. Wang, Q. Li, J. Wang, Y. Yan, Journal of Molecular Catalysis B:
Enzymatic 71 (2011) 45.
[122] Y. Liu, T. Liu, X. Wang, L. Xu, Y. Yan, Energy and Fuels 25 (2011) 1206.
[123] P. Hara, J.-P. Mikkola, D.Y. Murzin, L.T. Kanerva, Journal of Molecular Catalysis B:
Enzymatic 67 (2010) 129.
[124] P. Lozano, J.M. Bernal, G. Sanchez-Gomez, G. Lopez-Lopez, M. Vaultier, Energy &
Environmental Science 6 (2013) 1328.
[125] T. De Diego, A. Manjon, P. Lozano, M. Vaultier, J.L. Iborra, Green Chemistry 13
(2011) 444.
[126] Q. Pan, L. Yang, X. Meng, Journal of the American Oil Chemists Society (2013) 1.

You might also like