You are on page 1of 19

Biotechnology Advances 28 (2010) 500–518

Contents lists available at ScienceDirect

Biotechnology Advances
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / b i o t e c h a d v

Research review paper

Homogeneous, heterogeneous and enzymatic catalysis for transesterification of high


free fatty acid oil (waste cooking oil) to biodiesel: A review
Man Kee Lam, Keat Teong Lee ⁎, Abdul Rahman Mohamed
School of Chemical Engineering, Universiti Sains Malaysia, Engineering Campus, Seri Ampangan, 14300 Nibong Tebal, Pulau Pinang, Malaysia

a r t i c l e i n f o a b s t r a c t

Article history: In the last few years, biodiesel has emerged as one of the most potential renewable energy to replace current
Received 10 December 2009 petrol-derived diesel. It is a renewable, biodegradable and non-toxic fuel which can be easily produced
Received in revised form 16 March 2010 through transesterification reaction. However, current commercial usage of refined vegetable oils for
Accepted 20 March 2010
biodiesel production is impractical and uneconomical due to high feedstock cost and priority as food
Available online 31 March 2010
resources. Low-grade oil, typically waste cooking oil can be a better alternative; however, the high free fatty
Keywords:
acids (FFA) content in waste cooking oil has become the main drawback for this potential feedstock.
Biodiesel Therefore, this review paper is aimed to give an overview on the current status of biodiesel production and
Waste cooking oil the potential of waste cooking oil as an alternative feedstock. Advantages and limitations of using
Homogeneous catalyst homogeneous, heterogeneous and enzymatic transesterification on oil with high FFA (mostly waste cooking
Heterogeneous catalyst oil) are discussed in detail. It was found that using heterogeneous acid catalyst and enzyme are the best
Enzymatic option to produce biodiesel from oil with high FFA as compared to the current commercial homogeneous
Transesterification base-catalyzed process. However, these heterogeneous acid and enzyme catalyze system still suffers from
serious mass transfer limitation problems and therefore are not favorable for industrial application.
Nevertheless, towards the end of this review paper, a few latest technological developments that have the
potential to overcome the mass transfer limitation problem such as oscillatory flow reactor (OFR),
ultrasonication, microwave reactor and co-solvent are reviewed. With proper research focus and
development, waste cooking oil can indeed become the next ideal feedstock for biodiesel.
© 2010 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 501
2. Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 501
2.1. Composition of vegetable oils and fats . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
2.2. Composition of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
3. Current status of biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
4. Biodiesel costing and potential of waste cooking oil as feedstock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
5. Waste cooking oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
5.1. Thermolytic reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
5.2. Oxidative reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
5.3. Hydrolytic reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
6. Catalysis in transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 504
6.1. Homogeneous base-catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
6.2. Homogeneous acid-catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
6.3. Homogeneous acid and base-catalyzed transesterification: two steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 507
6.4. Heterogeneous base-catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 508
6.5. Heterogeneous acid-catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
6.5.1. Zirconium oxide (ZrO2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
6.5.2. Titanium oxide (TiO2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
6.5.3. Tin oxide (SnO2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 510

⁎ Corresponding author. Tel.: +60 4 5996467; fax: +60 4 5941013.


E-mail address: chktlee@eng.usm.my (K.T. Lee).

0734-9750/$ – see front matter © 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.biotechadv.2010.03.002
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 501

6.5.4. Zeolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511


6.5.5. Sulfonic ion-exchange resin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
6.5.6. Sulfonic modified mesostructure silica . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 511
6.5.7. Sulfonated carbon-based catalyst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
6.5.8. Heteropolyacids (HPAs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
6.6. Enzyme (biocatalyst) catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
6.6.1. Mucor miehei (Lipozym IM 60) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
6.6.2. Pseudomonas cepacia (PS 30) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
6.6.3. C. antarctica (Novozym 435). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
6.6.4. Bacillus subtilis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
6.6.5. Rhizopus oryzae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
6.6.6. Penicillium expansum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
7. Other methods or technologies for biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 514
7.1. Oscillatory flow reactor (OFR) for transesterification reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
7.2. Microwave technology in transesterification reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
7.3. Ultrasonic technology in transesterification reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
7.4. Co-solvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
8. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516

1. Introduction among other renewable resources as shown in Fig. 2 (International Energy


Agency I, 2008). Combustible renewable and waste accounted for 10.1%,
To date, fossil fuels account over 80.3% of the primary energy compared to hydro energy 2.2% and other 0.6% (included geothermal,
consumed in the world, and 57.7% of that amount is used in solar wind and heat). Hence, it is predicted that renewable energy from
transportation sector (International Energy Agency I, 2006). On the combustible energies such as biodiesel will enter the energy market
other hand, the global consumption of diesel fuel is estimated to be intensively in the near future to diversify the global energy sources.
934 million tonnes per year (Kulkarni and Dalai, 2006). Thus, The
World Energy Forum predicted that fossil oil will be exhausted in less 2. Biodiesel
than 10 decades, if new oil wells are not found (Sharma and Singh,
2009). The main reason that caused the fast diminishing of energy Biodiesel is an alternative diesel fuel derived from vegetable oils or
resources is due to rapid population and industrialization growth animal fats (Vasudevan and Briggs, 2008). The main components of
globally (Pimentel and Pimentel, 2006). Due to this phenomenon, the vegetable oils and animal fats are triglycerides or also known as esters
era of cheap crude oil no longer exists leading to high sky rocketing of fatty acids attached to a glycerol. Normally, triglycerides of
price of petroleum, bellicose conflicts and increasing the number of vegetable oils and animals fats consist of several different fatty
undernourished people especially from undeveloped countries. Fig. 1 acids. Different fatty acids have different physical and chemical
presents the projection of world energy demand in the near future properties and the composition of these fatty acids will be the most
indicating that there is an urgent need to find more new renewable important parameters influencing the corresponding properties of a
energies to assure energy security worldwide (Exxon Mobile, 2004). vegetable oils and animal fats (Gerhard Knothe and Krahl, 2004).
Renewable energy has been highlighted in the last ten years due to its Direct use of vegetable oils and animal fats as combustible fuel is not
potential to replace fossil fuel especially for transportation. Renewable suitable due to their high kinematic viscosity and low volatility.
energy sources such as solar energy, wind energy, hydro energy, and Furthermore, its long term use posed serious problems such as
energy from biomass and waste have been successfully developed and deposition, ring sticking and injector chocking in engine (Muniyappa
used by different nations to limit the use of fossil fuels. Nevertheless, based et al., 1996). Therefore, vegetable oils and animal fats must be subjected
on recent study from International Energy Agency (IEA), only energy
produced from renewable sources and waste has the highest potential

Fig. 2. World total energy supply by fuel (Mtoe) in year 2006 (excluding electricity and
Fig. 1. Projection of energy demand for the near future. heat tread). Total: 11,741 million tonnes of oil equivalent (Mtoe).
502 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

to chemical reaction such as transesterification to reduce the viscosity of Table 2


oils. In that reaction, triglycerides are converted into fatty acid methyl Chemical structures of common FAME.

ester (FAME), in the presence of short chain alcohol, such as methanol or Methyl ester Formula Common acronym Molecular weight
ethanol, and a catalyst, such as alkali or acid, with glycerol as a by-
Methy palmitic C17H34O2 C16:0 270.46
product (Vasudevan and Briggs, 2008). Eq. (1) depicts the transester- Methy stearate C19H38O2 C18:0 298.51
ification reaction. Another alternative way to produce biodiesel is Methy oleate C19H36O2 C18:1 296.50
through thermal cracking or pyrolysis. However, this process is rather Methy linoleate C19H34O2 C18:2 294.48
Methy linolenate C19H24O2 C18:3 292.46
complicated to operate and produce side products that have no
commercial value (Sharma and Singh, 2009).

3. Current status of biodiesel production

With the crude fossil fuel prices near all-time high, biodiesel has
emerged as the fastest growing industries worldwide. Several
countries especially United State and members of European Union
are actively supporting the production of biodiesel from the
agriculture sector. In year 2006, nearly 6.5 billion liters of biodiesel
was produced globally as shown in Fig. 3 (TBW, 2008). It is interesting
to note that 75% of the total biodiesel production comes from
European countries. This is mainly due to substantial support from the
ð1Þ European government such as consumption incentives (fuel tax
reduction) and production incentive (tax incentives and loan
2.1. Composition of vegetable oils and fats guarantees) which will further catalyzed the growth of the biodiesel
market in the next ten years. Besides that, the United States spent
Vegetable oils and animal fats usually have hydrophobic properties, around US$ 5.5 billion to 7.3 billion a year (TBW, 2008) to accelerate
which mean they are insoluble in water. As mention earlier, triglycerides biofuels production. As a result, around 450 million gallons of
are made up of 1 mol glycerol and 3 mol fatty acids. Fatty acids vary in biodiesel was produced in the United States in the year 2007
terms of carbon chain length and number of unsaturated bonds (double compared to only 25 million gallons in year 2004 (Thurmond,
bonds). Typical fatty acids compositions found in several vegetable oils 2008). This 1700% increment was indeed a shocking increase in the
are summarized in Table 1 (Ma and Hanna, 1999). Fatty acids that have entire history of biodiesel production.
no double bonds are termed “saturated” such as stearic acid. These However, by the year 2020, it is predicted that biodiesel
chains contain maximum number of possible hydrogen atoms per atom production from Brazil, China, India and some South East Asia
carbon. Fatty acids that have double bonds are termed “unsaturated” countries such as Malaysia and Indonesia could contribute as much
such as Linoleic acid. These chains do not contain maximum number of as 20% (Thurmond, 2008). The driving forces for development of
hydrogen atoms due to the presence of double bond(s) on some carbon biodiesel in these countries are economic, energy and environmental
atoms. Normally, natural vegetable oils and animal fats are obtained in security, improving trade balances and expansion of agriculture sector
the crude form through solvent extracting or mechanically pressing, (Zhou and Thomson, 2009). In addition, Brazil, China and India each
containing a lot of impurities such as free fatty acids, sterols and water. In have set targets to replace 5% to 20% of total diesel with biodiesel by
fact, these free fatty acids and water content will have significant effect the year 2010 with emphasis on second generation non-edible
on the transesterification reaction, especially if a base catalyst is used. feedstock (Thurmond, 2008). If governments from these countries
They could also interfere with the separation of FAME and glycerol continue to aggressively promote biodiesel generation and continue
during water washing (purification step) because of soap formation. to invest in research and development for non-edible feedstocks such
as jatropha, castor, algae and grease, the prospects to achieve
2.2. Composition of biodiesel biodiesel targets will be realized faster than anticipated. Fig. 4 depicts

Biodiesel is a mixture of fatty acid alkyl esters. In the case when


methanol is used as reactant, it will be a mixture of fatty acid methyl
esters (FAME) whereas if ethanol is used as reactant, the mixture will
be fatty acid ethyl esters (FAEE). However, methanol is commonly and
widely used in biodiesel production due to their low cost and
availability. Based on different feedstock, the biodiesel produced will
have different composition of FAME. Table 2 shows the common
composition of FAME in biodiesel (Ma and Hanna, 1999).

Table 1
Typical fatty acid composition (%) for different common oil source.

Fatty acid Soybean Cottonseed Palm Lard Tallow Coconut

Lauric (C12:0) 0.1 0.1 0.1 0.1 0.1 46.5


Myristic (C14:0) 0.1 0.7 1.0 1.4 0.8 19.2
Palmitic (C16:0) 0.2 20.1 42.8 23.6 23.3 9.8
Stearic (C18:0) 3.7 2.6 4.5 14.2 19.4 3.0
Oleic (C18:1) 22.8 19.2 40.5 44.2 42.4 6.9
Linoleic (C18:2) 53.7 55.2 10.1 10.7 10.7 2.2
Linolenic (C18:3) 8.6 0.6 0.2 0.4 0.4 0.0
Fig. 3. Biodiesel production in year 2006. Total production: 6.5 billion liters.
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 503

existence of governmental credits or subsidies. Based on Fig. 5, when


the feedstock cost is at US$ 0.52/kg (US$ 0.236/lb), the model
estimated a final biodiesel production cost of US$ 0.53/L ($ 2.00/gal).
It is clear that the cost of feedstock contributed the most, which
accounted 88% of the total production cost. In addition, the model also
estimated the process economic from the recovery of co-product,
glycerol as illustrated in Fig. 6. It was assumed that glycerol was sold
as low commercial grade glycerol with purity 80% w/w aqueous
solution. It can be noted that returns from selling glycerol was not
significant, only accounted to 6% reduction in the overall biodiesel
production cost.
In order to overcome this limitation, biodiesel manufacturer are
focusing their attention on using low-cost feedstock such as waste
cooking oil in order to ensure economic viability in biodiesel production.
Fig. 4. World biodiesel production and capacity. Waste cooking oil is far less expensive than refined vegetable oils and
therefore has become a promising alternative feedstock to produce
a more recent world biodiesel production and capacity in the recent biodiesel. In fact, generation of waste cooking oil in any country in the
years (Thurmond, 2008). world is huge, and may result to environmental contamination if no
proper disposal method is implemented. Table 3 shows the estimated
4. Biodiesel costing and potential of waste cooking oil as waste cooking oil produced in selected countries (Gui et al., 2008). Based
feedstock on the table, waste cooking oil generated is more than 15 million tonnes.
However, it should be noted that the actual amount of waste oil
Currently, the major concern for biodiesel production is economic produced is much higher based on global production. If such amount of
feasibility. In a reality scenario, biodiesel production will not be waste oil is successfully collected and converted to biodiesel, it will be
favored without tax exemption and subsidy from government; as the sufficient to meet the European biodiesel production target at 10 million
production cost is higher than fossil derived diesel (Demirbas and tonnes in year 2010 (Lam et al., 2009b).
Balat, 2006). The overall biodiesel cost consists of raw material Apart from that, a recent study on the production cost of biodiesel
(production and processing), catalyst, biodiesel processing (energy, using waste cooking oil as feedstock shows that the overall production
consumables and labour), transportation (raw materials and final costs of biodiesel can be reduced by more than half compared to virgin
products) and local and national taxes (Haas et al., 2006). To date, vegetable oil (Escobar et al., 2009). Furthermore, the production costs
most biodiesel plants are using refined vegetable oils as their main are even lower than fossil derived diesel as illustrated in Fig. 7
feedstock. Therefore, the cost of refined vegetable oils contributed (Escobar et al., 2009). Hence, the high cost of feedstock can be
nearly 80% of the overall biodiesel production cost (Lam et al., 2009b). overcome if waste cooking oil is used for biodiesel production.
Thus, it is undeniable that feedstock will be the most crucial variable
affecting the price of biodiesel in the global market. 5. Waste cooking oil
A generic process model to estimate biodiesel capital and
operating cost had been developed by Haas et al. and the result is Currently, world oil and fats production is about 154 million
shown in Fig. 5 (Haas et al., 2006). The model was designed on the tonnes (MPOC, 2008). This figure refer to the production of 17 major
basis of continuous transesterification process using crude, oils and fats, comprising from vegetable oils (i.e. soybean, cottonseed,
degummed soybean oil as the main feedstock and is dependent on groundnut, sunflower, rapeseed, sesame, corn, olive, palm, palm
the price of feedstock. In addition, the model was based on a process kernel, coconut, linseed, and castor) and animal fats/oils (i.e. butter,
plant with an annual production capacity of 378,541,181 L lard, tallow, grease and fish oil). Most of this oil is used for deep-frying
(10 × 106 gal). However, some economic factors were excluded, such processes, after which could cause disposal problem. Serious water
as internal rate of return, economic life span, corporate tax rate, contamination may occur if no proper disposal method is implemen-
salvage value, debt fracture, construction interest rate and long term ted. Such scenario does not only contribute to pollution problems but
interest rate, working capital, environment control equipment, is also harmful to human beings. In fact, EU has enforced a ban on the
marketing and distribution expenses, the cost of capital, and the utilization of all waste oils as domestic animal feed because during
frying process, many harmful compounds are formed. Eventually,

Fig. 5. Impact of feedstock prices on the predicted unit cost of biodiesel production from
crude degummed soybean oil with crude glycerol co-product assigned a value of $0.15/ Fig. 6. Impact of market value of 80% (w/w) glycerol on the unit cost of biodiesel
lb ($0.33/kg). production with the soy oil feedstock assigned a value of $0.236/lb ($0.520/kg).
504 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Table 3 5.2. Oxidative reaction


Quantity of waste cooking oil produced in selected countries.

Country Quantity (million tonnes/year) Oxidative reaction occurs when oxygen in air dissolved in the oil or
fat and reacts mainly with unsaturated acyglycerols (AG) resulting in
China 4.5
European 0.7–1.0 the formation of various oxidation products. The main reactions
United States 10.0 involved in the oxidation reactions are summarized in Fig. 8 (Velasco
Japan 0.45–0.57 and Dobarganes, 2002). RH represents triacylglycerol undergoing
Malaysia 0.5
oxidation in one of its unsaturated fatty acyl groups. Initially, radicals —
Canada 0.12
Taiwan 0.07 alkyl radicals (R˙) are formed. By the addition of oxygen, eventually
alkylperoxyl radicals (ROO˙) are produced. Finally, alkoxyl radicals
(RO˙) are formed due to the decomposition of hydroperoxides
these harmful compounds will enter the human food chain during (ROOH) which produce various saturated and unsaturated alde-
meat consumption (Kulkarni and Dalai, 2006). hydes, ketones, hydrocarbons, lactones, alcohols, acids and esters.
Since frying improves the taste of food, it has become a common Most of these compounds will remain within the oil or fat, e.g.
method in food preparation. During frying, oil is heated under dimeric and polymeric acid, dimeric AG and polyglycerols as prod-
atmospheric condition at temperature of 160–190 °C (Gazmuri and ucts of the radical reactions and increase the viscosity of the cooking
Bouchon, 2009) for relative long period of time. In addition, the same oil. Others might be further decomposed through alkoxyradicals to
oil or fat is used several times, mainly because of economical reasons. volatile polar compounds, e.g. hydroxyl- and epoxyacids that escape
However, continuously using the same oil or fat for frying will causes from the oil (Cvengros and Cvengrosova, 2004).
various physical and chemical changes in the oil, depending on the
type of oil and oil composition. Some physical changes observed in
5.3. Hydrolytic reaction
vegetable oil after frying are (i) an increase in viscosity, (ii) an
increase in specific heat, (iii) a change in surface tension, and (iv) a
The hydrolysis of triglycerides occurs when steam produced
change in color (Cvengros and Cvengrosova, 2004).
during the preparation of food. Part of the water quickly evaporates,
Apart from that, oils are also subjected to three types of reactions
but a certain part dissolved in the oil or fat and induces its cleavage to
during frying, mainly thermolytic, oxidative and hydrolytic (Mittel-
give higher fatty acids, glecerol, monoglycerides and diglycerides
bach and Enzelsberger, 1999; Nawar, 1984). These three reactions will
concentration (Kulkarni and Dalai, 2006).
continuously cause the formation of many undesired and harmful
compounds if the oil is used repeatedly. The toxicological effects of
these compounds upon human consumption are still not completely 6. Catalysis in transesterification
known. However, if waste cooking oil is to be made feedstock for
biodiesel production, the amount of polar compound in the waste The following section describes various catalysis methods for
cooking oil, especially free fatty acid (FFA) must be taken into transesterification reaction of waste cooking oil and the potential of
consideration as it will greatly affect the transesterification reaction. heterogeneous acid catalysts and enzymes towards a more sustainable
Refined oil usually contains less than 0.5 wt.% FFA whereas for waste biodiesel industry. However, due to limited availability of information
cooking oil, FFA contents range between 0.5 and 15 wt.% (Gerhard regarding the transesterification of waste cooking oil for certain type of
Knothe and Krahl, 2004). catalysts; other oils with high FFA oils are taken to re-present waste
cooking oil so that the potential of these catalysts for transesterification
can be easier understood. Nonetheless, study that uses waste cooking oil
5.1. Thermolytic reaction as feedstock will be given priority when discussing various technologies.
Table 4 summarizes the reaction conditions for various types of catalysts
A thermolytic reaction occurs in the absence of oxygen at high
temperatures. A series of alkanes, alkenes, lower fatty acids,
symmetric ketones, oxopropyl esters, CO, and CO2 are produced
from the saturated fatty acids in the oil. For unsaturated fatty acids,
basically diametric compounds including dehydrodimers, saturated
dimers and polycyclic compounds are formed. In addition, dimers and
trimers may be formed when unsaturated fatty acids react with other
unsaturated fatty acids through Diels–Alder reaction (Kulkarni and
Dalai, 2006).

Fig. 7. Range of production costs for biodiesel and diesel in year of 2006. Fig. 8. Simplified mechanism for oil oxidation reaction during frying.
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 505

Table 4
Comparison of reaction conditions and performance for various types of catalysts used in transesterification of waste cooking oil.

Catalyst Reaction conditions Performance Comments Reference

Temperature, °C Type of alcohol Catalyst Reaction


(alcohol to oil loading, time, h
molar ratio) wt.%

Homogeneous base catalyst


NaOH 60 Methanol (7:1) 1.1 0.33 Yield = 88.8% Excess catalyst used Leung and Guo
led to soap formation (2006)
KOH 87 Methanol (9:1)a 6 2 Yield = 87%a – Demirbas (2009)

Homogeneous acid catalyst


H2SO4 95 Methanol (20:1) 4 20 Conversion = N 90% – Wang et al. (2006)
H2SO4 70 Methanol (245:1) 41.8a 4 Yield = 99% – Zheng et al. (2006)
H2SO4 65 Methanol (30:1) 1 69 Conversion = 99% – Freedman et al.
(1984)

Two-step: Acid catalyst follow by base catalyst


Ferric sulfate follow Acid: 95 Acid: Methanol (10:1) Acid: 2 Acid: 2 Conversion = 97% – Wang et al. (2006)
by KOH Base: 65 Base: Methanol (6:1) Base: 1 Base: 1
Ferric sulfate follow Acid: 100 Acid: Methanol (9:1) Acid: 2 Acid: 2 Yield = 96% – Patil et al. (2010)
by KOH Base: 100 Base: Methanol (7.5:1)a Base: 0.5a Base: 1
Ferric sulfate follow Acid: 60 Acid: Methanol (7:1) Acid: 0.4 Acid: 3 Yield = 81.3% – Wan Omar et al.
by CaO (2009)
Base: 60 Base: Methanol (7:1) Base: Not Base: 3
clearly
specified

Heterogeneous base catalyst


CaO Reflux methanol Methanol (12:1) 0.85 1 Yield = 66% Calcium soap was Kouzu et al. (2008a)
temperature: 60–65 detected
K3PO4 60 Methanol (6:1) 4 2 Yield = 97.3% – Guan et al., 2009
Oil palm ash 60 Methanol (18:1) 5.35 0.5 Yield = 71.7% – Chin et al. (2009)

Heterogeneous acid catalyst


WO3/ZrO2 75 Methanol to Oleic acid Not clearly 20 FFA Packed-bed reactor Park et al. (2008)
(19.4:1) specified Conversion = 85% was used
Zeolite Y (Y756) 460 Methanol (6:1) Not clearly 0.37 Yield = 26.6% – Brito et al. (2007)
specified
Carbon-based catalyst 80 Methanol (30:1) 10 8 Yield = 92% – Lou et al. (2008)
derived from starch
H3PW12O40·6H2O 65 Methanol (70:1) 3.7a 14 Yield = 87% 4 Å zeolite was used Cao et al. (2008)
(PW12) simultaneously
to adsorb water
Zr0.7H0.2PW12O40 65 Methanol (20:1) 2.1 8 Yield = 98.9% Nanostructured Zhang et al. (2009)
(ZrHPW) catalyst contain
double acid sites
ZS/Si 200 Methanol (18:1) 3 5 Yield = 98% – Jacobson et al.
(2008)
SO2−
4 /TiO2–SiO2 200 Methanol (9:1) 3 4 a
Yield = 90% a
– Peng et al. (2008)
SO2−
4 /SnO2–SiO2 150 Methanol (15:1) 3 3 Yield = 92.3% – Lam et al. (2009a)

Enzymatic catalyst
Pseudomonas cepacia 38.4 Ethanol (6.6:1) 13.7 2.47 Yield = 96% Second portion of lipase Wu et al. (1999)
(PS 30) SP 435 (5 wt %) was
added into the reaction
media after 1 h
Candida antarctica 30 Methanol (3:1) 4 50 Conversion = 90.4% Three-step batch Watanabe et al.
(Novozym 435) methanolysis (2001)
Novozym 435 30 Methanol (3:1) 4 50 Conversion = 90.9% Three-step continuous Watanabe et al.
methanolysis at different (2001)
flow rate
Novozym 435 40 Methanol (4:1) 4 12 Yield = 88 Tert-butanol was used as Halim and Harun
co-solvent Kamaruddin (2008)
Bacillus subtilis 40 Methanol (1:1) 3 72 Yield = 90 Methanol was added in Ying and Chen
encapsulated in two stepwise (2007)
magnetic particles
(Magnetic cell
biocatalyst)
Rhizopus oryzae 40 Methanol (4:1) 30 30 Yield = 88–90% Methanol was added in Chen et al. (2006)
three stepwise.
Immobilized Penicillium 35 Methanol (1:1) Not clearly 7 Yield = 92.8% Methanol was added in Li et al. (2009)
expansum on resin specified three stepwise.
D4020 Blue silica gel was used
to adsorb excess water.
a
Self-estimation.
506 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Table 5
Advantages and disadvantages of different types of catalysts used in transesterification of waste cooking oil.

Type of catalyst Advantages Disadvantages

Homogeneous base catalyst •Very fast reaction rate — 4000 times faster than acid-catalyzed •Sensitive to FFA content in the oil
transesterification
•Reaction can occur at mild reaction condition •Soap will formed if the FFA content in the oil is
and less energy intensive more than 2 wt.%
•Catalysts such as NaOH and KOH are relatively •Too much soap formation will decrease the biodiesel yield and
cheap and widely available cause problem during product purification especially generating
huge amount of wastewater
Heterogeneous base catalyst •Relatively faster reaction rate than acid-catalyzed •Poisoning of the catalyst when exposed to ambient air
transesterification
•Reaction can occur at mild reaction condition and •Sensitive to FFA content in the oil due to its basicity property
less energy intensive
•Easy separation of catalyst from product •Soap will be formed if the FFA content in the oil is more
than 2 wt.%
•High possibility to reuse and regenerate the catalyst •Too much soap formation will decrease the
biodiesel yield and cause problem during product
purification
•Leaching of catalyst active sites may result to product
contamination
Homogeneous acid catalyst •Insensitive to FFA and water content in the oil •Very slow reaction rate
•Preferred-method if low-grade oil is used •Corrosive catalyst such as H2SO4 used can lead to corrosion on
reactor and pipelines
•Esterification and transesterification occur simultaneously •Separation of catalyst from product is problematic
Reaction can occur at mild reaction condition and less
energy intensive
Heterogeneous acid catalyst •Insensitive to FFA and water content in the oil •Complicated catalyst synthesis procedures lead to higher cost
•Preferred-method if low-grade oil is used •Normally, high reaction temperature, high alcohol to oil molar
ratio and long reaction time are required.
•Esterification and transesterification occur simultaneously •Energy intensive
•Easy separation of catalyst from product •Leaching of catalyst active sites may result to product
contamination
•High possibility to reuse and regenerate the catalyst
Enzyme •Insensitive to FFA and water content in the oil •Very slow reaction rate, even slower than acid-catalyzed
transesterification
•Preferred-method if low-grade oil is used •High cost
•Transesterification can be carried out at low reaction •Sensitive to alcohol, typically methanol that can deactivate
temperature, even lower than homogeneous base catalyst the enzyme
•Only simple purification step is required

in transesterification of waste cooking oil whereas Table 5 lists out their and base catalyst (KOH). This reaction is highly undesirable because it
advantages and disadvantages. will deactivate the catalyst from accelerating the transesterification
reaction. Furthermore, excessive soap in the products can drastically
6.1. Homogeneous base-catalyzed transesterification reduce the fatty acid methyl ester (FAME) yield and inhibit the
subsequent purification process of biodiesel, including glycerol
Currently, biodiesel is commonly produced using homogeneous separation and water washing (Nag, 2008, Kulkarni and Dalai, 2006).
base catalyst, such as sodium hydroxide (NaOH) or potassium
hydroxide (KOH) (Felizardo et al., 2006, Kulkarni and Dalai, 2006).
These catalysts are commonly used in the industries due to several
reasons: (i) able to catalyze reaction at low reaction temperature and
atmospheric pressure; (ii) high conversion can be achieved in a
minimal time, (iii) widely available and economical (Lotero et al., ð2Þ
2005). In fact, it was reported that the rate for base-catalyzed reaction
would be 4000 times faster compared to acidic catalyst (Fukuda et al., Apart from that, high water content in waste cooking oil also
2001, Kulkarni and Dalai, 2006). However, the use of this catalyst is affects the methyl ester yield. When water is present, particularly at
limited only for refined vegetable oil with less than 0.5 wt.% FFA high temperatures, it can hydrolyze triglycerides to diglycerides and
(Wang et al., 2006) or acid value less than 1 mg KOH/g (Felizardo et form free fatty acid. Eq. (3) shows the hydrolysis reaction. With the
al., 2006). Some researchers reported that base catalyst can tolerate presence of base catalyst, the FFA will subsequently react to form soap
higher content of FFA as shown in Table 6. Nevertheless, it is clear that as shown in Eq. (2). Thus, when water is present in the reactant, it
the FFA content in oil feedstock should be as low as possible (ranging
from less than 0.5 wt.% to less than 2 wt.%) for base-catalyzed
transesterification reaction. Thus, if waste cooking oil with an average Table 6
Level of FFA recommended for homogeneous base catalyst transesterification.
FFA content more than 6 wt.%, base catalyst is definitely not suitable
to be used (Lotero et al., 2005). Author/reference Recommended FFA (wt.%)
FFA consists of long carbon chain that is disconnected from Ma and Hanna (1999) b1
glycerol backbone. They are sometimes called carboxylic acids. If an Ramadhas et al. (2005) ≤2
oil or fat containing high FFA such as oleic acid is used to produce Zhang et al. (2003a) b 0.5
biodiesel, alkali catalyst will typically react with FFA to form soap, Freedman et al. (1984) b1
Kumar Tiwari et al. (2007) b1
which is highly undesirable (Nag, 2008, Yan et al., 2009, Kulkarni and
Sahoo et al. (2007) ≤2
Dalai, 2006). Eq. (2) shows a typical reaction between FFA (oleic acid)
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 507

generally manifests itself through excessive soap production. Apart


from that, the soaps of saturated fatty acids tend to solidify at ambient
temperatures and thus a reaction mixture with excessive soap may
gel-up and form a semi-solid mass which is very difficult to recover
(Felizardo et al., 2006).

ð3Þ

6.2. Homogeneous acid-catalyzed transesterification

Since liquid base-catalyzed transesterification process poses a lot


of problems especially for oil or fat with high FFAs concentration,
liquid acid catalysts are proposed in order to overcome the
limitations. To date, the most investigated catalysts for acid-catalyzed
system are sulfuric acid (H2SO4) and hydrochloric acid (HCl). Acid-
catalyzed transesterification holds an important advantage with
respect to base-catalyzed process: acid catalyst is insensitive to the
presence of FFAs in the feedstock (Kulkarni and Dalai, 2006) and can
catalyzes esterification and transesterification simultaneously (Jacob-
son et al., 2008). Esterification is a chemical reaction in which two
reactants, typically an alcohol (e.g. methanol) and an acid (e.g. FFA)
react to form an ester as the reaction product. It was reported that acid
Fig. 9. Homogeneous acid-catalyzed reaction mechanism for the transesterification of
catalysis is more efficient when the amount of FFA in the oil exceeds
triglycerides: (1) protonation of the carbonyl group by the acid catalyst; (2)
1 wt.% (Zhang et al., 2003a, Canakci and Van Gerpen, 1999, Freedman nucleophilic attraction of the alcohol, forming a tetrahedral intermediate; (3) proton
et al., 1984). In addition, economic analysis has proven that acid- migration and breakdown of the intermediate. The sequence is repeated twice for R2
catalyzed procedure, being a one-step process, is more economical and R3.
than the base-catalyzed process which requires an extra step to
convert FFA to methyl esters (Zhang et al., 2003a, b). pathway (formation of electrophilic species by acid catalysis versus
However, acid-catalyzed system is not a popular choice for formation of stronger nucleophile by base catalysis) is ultimately
commercial applications due to slower reaction rate, requirement of responsible to the difference in catalytic activity in the transesterifica-
high reaction temperature, high molar ratio of alcohol to oil, tion reaction.
separation of the catalyst, serious environmental and corrosion
related problem (Jacobson et al., 2008, Wang et al., 2006). In a study
of acid-catalyzed transesterification of waste cooking oil using H2SO4, 6.3. Homogeneous acid and base-catalyzed transesterification: two steps
Wang et al. reported that the yield of FAME increased with longer
reaction time, higher methanol to oil ratio and higher catalyst loading. Since homogeneous acid and base catalysts have their own
The conversion of waste cooking oil was more than 90% at a reaction advantages and limitations, some studies have attempted to use a
time of 10 h with ratio of methanol to oil at 20:1 and 4 wt.% H2SO4 combination of both catalysts to synthesis biodiesel from oil contain-
(with reference to weight of oil) (Wang et al., 2006). In another study, ing high FFAs. Initially, acid catalyst was employed to convert FFAs to
Freedman et al. reported 99% oil conversion by using 1 mol% of H2SO4 ester through esterification. When the FFAs content in the oil drops to
and methanol to oil ratio 30:1 for 69 h reaction time (Freedman et al., lower than 0.5–1 wt.%, transesterification of the oil can then be
1984). These data indicates that acid-catalyzed transesterification performed by using a base catalyst.
process requires more severe reaction conditions (such as longer Canakci and Van Cerpen have developed a pilot plant to produce
reaction time) than base-catalyzed reaction. biodiesel from feedstock with high FFAs content via two steps
Lotero et al. investigated the reasons for the low activity of acid- method; esterification and transesterification (Canakci and Van
catalyzed compared to base-catalyzed transesterification reaction. Fig. 9 Gerpen, 2003). The feedstock was first treated with H2SO4 to reduce
shows the mechanism of acid-catalyzed reaction (Lotero et al., 2005). the level of FFAs to below 1 wt.%, followed by transesterification
From the figure, it can be seen that the protonation of carbonyl group is process catalyzed by homogeneous base KOH. Although high FAME
the key step in the catalyst–reactant interaction. However, this initial yield can be obtained, but the rate of FFA esterification reaction was
chemical pathway has in turn increases the electrophilicity of the relatively very slow. Thus, higher amount of acid catalyst are required
adjoining carbon atom, resulting in the intermediate molecules to accelerate the rate of reaction. The drawback of this two-step
susceptible to nucleophilic attack. In contrast, the base catalysis takes process is even more pronounced due to the requirement of extra
on a more direct route, at which alkoxide ion is created initially and separation steps to remove the catalyst in both stages. Although
directly acts as a strong nucleophile. Fig. 10 shows the mechanism of problem of catalyst removal from the first stage can be avoided by
base-catalyzed reaction (Lotero et al., 2005). This crucial different using base catalyst from the second stage through neutralization
508 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Fig. 11. CaO as heterogeneous base catalyst mechanism in transesterification of


triglycerides: (1) abstraction of proton from methanol by the basic sites to form
methoxide anion; (2) methoxide anion attacks carbonyl carbon in a molecule of the
triglyceride leading to the formation of alkoxycarbonyl intermediate; (3) alkoxycarbo-
Fig. 10. Homogeneous base-catalyzed mechanism for the transesterification of triglycer-
nyl intermediate further transformed into a more stable form: FAME and anion of
ides: (1) production of the active species, RO−; (2) nucleophilic attack of RO− to carbonyl
diglyceride; (4) methoxide cation attracts the anion of diglyceride leading to the
group on triglycerides, forming a tetrahedral intermediate; (3) intermediate breakdown;
formation of diglyceride. The sequence is repeated twice for R2 and R3.
(4) regeneration of the RO− active species. The sequence is repeated twice for R2 and R3.

process, the use of extra base catalyst will add to the cost of biodiesel 3065 ppm which exceeded the basic standard of biodiesel, the
production (Kulkarni and Dalai, 2006). concentration of mineral matter should be below 200 ppm (Kouzu
et al., 2008a).
6.4. Heterogeneous base-catalyzed transesterification In addition, some researchers also pointed out that soluble
substance from CaO can leached out during transesterification.
To date, many solid base catalysts have been developed for Gryglewicz (1999) stated in his paper that calcium oxide slightly
biodiesel production, such as basic zeolites, alkaline earth metal dissolves in methanol. Lopez et al. conducted transesterification of
oxides and hydrotalcites. On top of that, alkaline earth metal oxides sunflower oil with methanol in which a slight amount of calcium
especially calcium oxide, CaO have attracted much attention due to oxide was found to dissolved in the reaction product (Granados et al.,
their relatively high basic strength, low solubility in methanol and can 2007). Kouzu et al. (2008b) further identified the soluble substance as
be synthesized from cheap sources like limestone and calcium calcium diglyceroxide in which CaO reacted with glycerol during
hydroxide (Zabeti et al., 2009). Kouzu et al. (2008a) reported that transesterification of soybean oil with methanol. Thus, an extra
CaO obtained from calcinations of pulverized limestone, CaCO3 at purification step is needed such as ion-exchange resin to remove the
900 °C for 1.5 h in the flow of helium gas exhibited substantially good soluble content in the biodiesel (Kouzu et al., 2009).
result in transesterification of refined soybean oil. The yield of FAME Apart from that, Granados et al. (2007) used activated CaO as a
was 93% after 1 h reaction time at methanol reflux temperature and solid base catalyst in the transesterification of sunflower oil to
methanol to oil ratio 12:1. Fig. 11 shows the mechanism of CaO as investigate the role of water and carbon dioxide on the deterioration
heterogeneous base catalyst in transesterification. However, the yield of the catalytic performance upon contact with air for different period
of FAME dropped to 66% when waste cooking oil with FFA content of time. The study showed that CaO was rapidly hydrated and
2.6 wt.% was used under the same reaction condition. It is obvious that carbonated in the air. No calcium oxide peak was detected in the
the basic sites of CaO were poisoned by strong adsorption of FFAs on samples after exposed to air for more than 20 days. It was further
the surface of the catalyst (Kouzu et al., 2008a). Consequently, a reported that the active sites of CaO were poisoned due to
portion of the catalyst changed into calcium soap by reacting with the chemisorption of carbon dioxide and water on the surface sites to
FFAs adsorbed, resulting in low recovery of catalyst. Kouzu et al. form carbonates and hydroxyl groups, respectively. However, the
further reported that the concentration of Ca in reaction product was catalytic activity of CaO can be regenerated if CaO is subjected to an
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 509

activation treatment at 700 °C in order to remove the main poisoning transesterification of palm kernel oil and crude coconut oil with
species (the carbonate groups) from the surface. However, leaching of methyl ester yield reaching as high as 90.3% and 86.3%, respectively.
the catalyst was still observed in the transesterification reaction However, when unsulfated ZrO2 was used as catalyst instead of SO2− 4 /
although prior thermal treatment was employed. ZrO2, only 64.5% (palm kernel oil) and 49.3% (crude coconut oil) of
Besides, magnesium oxide (MgO) which is produced by direct methyl ester yield were attained, respectively. This eventually
heating of magnesium carbonate or magnesium hydroxide was also indicates that modification of metal oxide surface acidity is the key
investigated on its catalytic activity in transesterification. Di Serio et al. factor in obtaining high conversion of triglycerides.
(2006) reported that MgO was efficient in transesterification of soybean Apart from that, combination of alumina, Al2O3 with ZrO2 and
oil, however high reaction temperature (180 °C) is required. At low modification of ZrO2–Al2O3 with tungsten oxide (WO3) not only
reaction temperature (100 °C), MgO catalyst exhibited very low provides high mechanical strength but also enhances the acidity of the
catalytic performance since the FAME yield observed was less than catalyst (Jacobson et al., 2008). The addition of Al2O3 further stabilizes
20%. This result was in agreement with Cantrell et al. (2005) and the tetragonal phase of ZrO2 support and also prevents the growth of
Gryglewicz (1999), who reported low or even no activity when using WO3 particles. Furuta et al. (2004a) evaluated the performance of
MgO in transesterification reactions performed at 60 °C. In fact, MgO has tungstated zirconia–alumina (WZA) and sulfated zirconia–alumina
the weakest basic strength among group II oxides, e.g. CaO and (SZA) in the transesterification of soybean oil with methanol at 200–
strontium oxide (SrO) (Kouzu et al., 2008a). Nevertheless, mixed 300 °C using a fixed bed reactor under atmospheric pressure. WZA
magnesium–alumina (Mg–Al) oxide prepared by using hydrotalcites was found to have a higher activity in transesterification as compared
(Mg6Al2(OH)16CO34H2O) as precursor and calcined at high temperature to SZA. However, the authors did not elaborate much on the causes of
has manifested the basic sites of the catalyst (Xie et al., 2006). More than the improved activity of WZA catalyst. Nevertheless, high reaction
90% of FAME yield was observed when using Mg–Al oxide as catalyst but temperature (250 °C) and long reaction time (20 h) were needed in
high reaction temperature is still required (Di Serio et al., 2006). Besides order to achieve 90% conversion. Similar results was also reported by
that, high concentration of Mg and Al ions were found to leach out from Jacobson et al. where by catalyst was prepared by impregnating 10 wt.
the catalyst, resulting in the requirement of extra purification step (Oku % of WO3 on to ZrO2–Al2O3. Relatively low ester yield was obtained
et al., 2005). However, a more in-depth study on using Mg–Al in (65%) when the transesterification reaction was carried out at lower
transesterification of high FFAs oil is required since the effect of FFAs reaction temperature of 200 °C and shorter reaction time of 10 h
content towards the catalyst performance is still unknown. (Jacobson et al., 2008). On the other hand, Faria et al. (2009) proposed
a reaction mechanism for ZrO2 supported on SiO2 in transesterifica-
6.5. Heterogeneous acid-catalyzed transesterification tion of soybean oil, as shown in Fig. 12. Thus, a more detail
understanding on reaction pathway by ZrO2 catalyst in transester-
Currently, biodiesel research is focused on exploring new and ification can be obtained.
sustainable solid acid catalysts for transesterification reaction. In Application of ZrO2 in esterification of waste cooking oil was also
addition, it is believed that solid acid catalysts have the strong reported by Park et al. (2008). In their study, WO3 was incorporated
potential to replace liquid acid catalyst (Jacobson et al., 2008). The into ZrO2 rather than impregnating with H2SO4. It was found that
advantages of using solid acid catalyst are (1) they are insensitive to WO3/ZrO2 has higher stability than SO2− 4 /ZrO2, and therefore avoiding
FFA content, (2) esterification and transesterification occurs simulta- the leaching of acid sites into the reaction media. Even if WO3 leached
neously (Kulkarni and Dalai, 2006), (3) eliminate the washing step of into the reaction media, it does not contaminate the product (Park et
biodiesel (Jitputti et al., 2006), (4) easy separation of the catalyst from al., 2008, Park et al., 2010). From the study, it was found that 85% of
the reaction medium, resulting in lower product contamination level, FFA conversion was attained in a packed-bed reactor after 20 h of
(5) easy regeneration and recycling of catalyst and (6) reduce reaction time at 75 °C, but decrease to 65% and remained stable
corrosion problem, even with the presence of acid species (Suarez thereafter. The reason given was due to the oxidation of WO3 after
et al., 2007). In fact, the development of heterogeneous catalyst long term exposure to FFA (a reducing agent) that resulted to a
system holds an important factor to be incorporated into a continuous decrease in catalytic activity. Therefore, leaching of WO3 was rule out
flow reactor (Lotero et al., 2005). Such continuous process can as the main reason for catalyst deactivation. In addition, WO3/ZrO2
minimize product separation and purification costs, making it could be simply regenerated by air re-calcination. However, further
economically viable and able to compete with commercial petro- study on catalyst and process optimization and also the oxidation
leum-based diesel fuel (de Almeida et al., 2008). state of WO3 are still required.
The ideal solid acid catalyst for transesterification reaction should Despite the high acidity of SO2− 4 /ZrO2, however it is known to
have characteristics such as an interconnected system of large pores, a suffer significant deactivation during liquid-phase transesterification,
moderate to high concentration of strong acid sites, and a hydropho- possibly due to sulfate leaching (Omota et al., 2003). Catalyst leaching
bic surface (Kulkarni and Dalai, 2006). However, research on direct was tested by dissolving fresh SO2− 4 /ZrO2 catalyst with water. The pH
use of solid acid catalyst for biodiesel production has not been widely of the suspension was found to decrease quickly as a result of sulfate
explored due to its limitation of slow reaction rate and possible groups hydrolyzed to become H2SO4 and HSO− 4 . This will then cause
undesirable side reactions. Furthermore, there is a knowledge gap on the transesterification to occur via homogeneous acid catalysis and
the fundamental studies dealing with reaction pathway of triglycer- therefore interfere with the measurements of heterogeneous catalytic
ides on solid acids. The following section will give an overview of activity. Recently, a new preparation method for SO2− 4 /ZrO2 was
various solid acid catalysts reported so far for biodiesel production. proposed by using chlorosulfonic acid, HSO3Cl instead of the
conventional impregnation of H2SO4 (Yadav and Murkute, 2004).
6.5.1. Zirconium oxide (ZrO2) The prepared SO2− 4 /ZrO2 exhibited higher catalytic activity in
There have been several studies on the usage of zirconium oxide transesterification and no sulfate leaching was observed. However,
(ZrO2) as a solid acid catalyst for transesterification of different HSO3Cl is a very hazardous chemicals, even an exposure for a very
feedstock due to its strong surface acidity. The acidity property can short time can cause death or major fatal injury (Kapias and Griffiths,
even be enhanced by coating the surface of this metal oxide with 2001).
anions like sulfate and tungstate. This can be done by impregnating
ZrO2 with acidic solution such as sulfuric acid (H2SO4) to become 6.5.2. Titanium oxide (TiO2)
sulfated zirconia, SO2−
4 /ZrO2 (Miao and Gao, 1997). Jitputti et al. Titanium dioxide (TiO2) is among transition metal oxides that
(2006) reported that SO2− 4 /ZrO2 can gives promising results in have attracted attention for biodiesel production due to their acidic
510 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Fig. 12. Proposed catalytic cycle for SiO2/ZrO2 in transesterification.

properties. In addition, introduction of sulfuric group on the surface of properties make SnO2 an excellent material for several applications
TiO2 will enhance the acid strength of the catalyst. However, there are in catalysis, conductivity, gas sensing, ceramics, plastics and biomed-
still very limited study reported in the literature describing the icine (Toledo-Antonio et al., 2003). Normally, SnO2 with mesostruc-
application of such catalyst for transesterification of vegetable oils (de ture form can be obtained by hydrolyzing inorganic precursor such as
Almeida et al., 2008). Recently, Chen et al. (2007) had evaluated the meta-stannic acid (SnO·H2O) (Furuta et al., 2004a) in the presence of
catalytic activity of SO2− 2−
4 /TiO2 and SO4 /ZrO2 for transesterification of cationic, anionic and neutral surfactants as structure director
cotton seed oil high in FFAs content to FAME. It was interesting to find (Gutierrez-Baez et al., 2004). However, most of SnO2 mesostructure
that the activity of this catalyst is proportional to its specific surface is not stable and collapsed during calcination. Thus, in order to
area. SO2− 2
4 /TiO2 with a specific surface area of 99.5 m /g can achieved increase the mesostructure stability, anionic surfactant with phos-
a higher yield of 90% as compared to SO2− 4 /ZrO2 with a specific surface phate or sulfate has been proposed such as the synthesis of
area of 91.5 m2/g, which can only achieved a yield of 85%. However, mesoporous sulfated tin oxide, SO2− 4 /SnO2 (Gutierrez-Baez et al.,
this catalyst requires high reaction temperature (230 °C), a negative 2004). The phosphate or sulfate groups will then remain bounded to
factor for industrial application. When lower reaction temperature the surface of SnO2, which will further stabilized its mesostructure
(120 °C) and reaction time (1 h) were used, the FAME yield obtained walls and increasing its thermal stability.
was only 40% (de Almeida et al., 2008). Apart from SO2− 2−
4 /ZrO2, SO4 /SnO2 is another solid super acid
It was proposed that the reactivity of SO2−4 /TiO2 can be increased by catalyst that has potential in transesterification reaction. Results from
introducing a secondary metal, SiO2 to produce SO2− 4 /TiO2–SiO2 (Peng temperature programmed desorption (TPD) of ammonia and adsorp-
et al., 2008). By adding SiO2 to SO2−
4 /TiO2, the specific surface area of the tion heat of argon (Ar) indicates that the acid strength of SO2− 4 /SnO2
catalyst increased to 258 m2/g with an average pore diameter of was higher than that of SO2− 2−
4 /ZrO2. Consequently, SO4 /SnO2 have
10.8 nm. The synthesized SO2− 4 /TiO2–SiO2 was then subjected to shown superior catalytic activity than SO2− 4 /ZrO2 in esterification of n-
transesterification of refined cotton seed oil blended with 50% oleic octanoic acid with methanol at temperature below 150 °C (Furuta et
acid. The optimum FAME yield obtained was more than 90% at reaction al., 2004b). However, report concerning the application of SO2− 4 /SnO2
temperature 200 °C, methanol to oil ratio 9:1, catalyst loading 3 wt.% in transesterification reaction is still very limited, typically for oil or fat
and a shorter reaction time of 3 h. However, the reaction temperature is containing high FFA. In addition, among the few studies reported in
still considered very high if compared to homogeneous catalyst which the literature, the findings reported on its catalytic performance in
lies between 60 and 100 °C. Apart from that, more in-depth research transesterification reaction have not been conclusive. The limited
especially on the leaching characteristic of this catalyst must be carried study carried out on SO2− 4 /SnO2 may be due to its complicated
out to justify its potential for industrial biodiesel production. preparation procedure particularly owing to the difficulty in obtaining
oxide gels from its salt as compared to SO2− 4 /ZrO2 that can be easily
6.5.3. Tin oxide (SnO2) prepared (Khder et al., 2008, Matsuhashi et al., 2001). Therefore, a
Tin oxide, SnO2 is n-type semiconductor material with a wide band more widespread testing of SO2− 4 /SnO2 should be carried out to obtain
gap (∼ 3.6 eV); it is transparent to visible light and reflects infrared more data and information for this promising material as potential
light (Gutierrez-Baez et al., 2004). This optical and electrical solid catalyst for transesterification reaction.
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 511

6.5.4. Zeolites found to require the least reaction time to achieve 90% conversion
Zeolites are microporous crystalline solids which contain silicon compared to sulfated zirconia and Nafion-NR50. Similar good results
(Si), aluminium (Al) and oxygen in their framework. One of the in esterification by using acidic Amberlyst-15 were also reported by
common applications of zeolite (inorganic solid catalyst) is for the Heidekum et al. (1999) and Chen et al. (1999). However, Amberlyst-
production of organic compounds such as ester (Balaji and Chanda, 15 was found to give low performance in transesterification reaction.
1998). This is because the characteristic of zeolite can be tailored At a relative low reaction temperature (60 °C), the conversion of
made to suit its function. For instance, zeolite can be synthesized with sunflower oil to FAME was reported to be only 0.7%, using the
different crystal structures, pore sizes, Si/Al ratios and proton- following reaction conditions: atmospheric pressure for 8 h reaction
exchange levels (Lotero et al., 2005). Therefore, the acid strength of time and 6:1 methanol to oil molar ratio (Vicente et al., 1998). In
the catalyst can be controlled by changing the aluminosilicate another study, Dos Reis et al. reported the transesterification of
framework such that it fits specific reaction requirements (Corma Babassu coconut oil using Amberlyst-15 (Dos Reis et al., 2005). It was
and Garcia, 1997). However, it should be noted that low surface reported that a rather good triglycerides conversion of 80% can be
acidity may caused slow reaction rate whereas extremely high acidity achieved only if the methanol to oil ratio used is increased to 100:1.
may cause deactivation due to coking or possible formation of The reaction temperature and time are at 60 °C and 8 h respectively.
undesirable by-products. Apart from that, zeolite also has another key This finding was rather expected as the activity of solid acid catalysts
feature. Since zeolite with specific pore structure and surface in transesterification is normally low at low reaction temperature.
hydrophobicity can be tailored made according to substrate's size Consequently, if Amberlyst-15 is to be used, it is necessary to increase
and polarity (Lotero et al., 2005), therefore only molecules with the reaction temperature to 150–200 °C to obtain sufficiently fast
appropriate dimension are allowed to enter the zeolite cavity and reaction rate. However, most ion-exchange resins such as Amberlyst-
diffusing through the pores (Xavier et al., 2009). 15 have low thermal stability and become unstable at temperature
Although zeolite has numerous advantages over other heteroge- above 140 °C (Lotero et al., 2005). Thus, this problem certainly limits
neous catalysts, however its catalytic activity in transesterification their application to reactions that require high temperatures. Study
reactions is relatively low. This is mainly due to diffusion limitation of regarding the deactivation of polystyrene sulfonic acid resins in
bulky reactants (triglycerides) into the microporous structure of esterification of high FFAs oils at a higher reaction temperature was
zeolite (Kiss et al., 2006). Triglycerides with an average molecular size reported and discussed extensively by Tesser et al. (2005).
of 2 nm suffered mass transfer resistance in the zeolite's micropore
(1–2 nm). Therefore, it is believed that transesterification reaction 6.5.6. Sulfonic modified mesostructure silica
only occurs on the external surface of zeolite crystal. In order to Mesostructure materials such as silica have an exceptional
overcome this limitation, the pore size and structure of zeolite must potential to be utilized as heterogeneous acid catalyst in biodiesel
be adjusted by varying the Si/Al ratio. Generally, higher Si/Al ratio production. This mesoporous materials (silica) consist of large
resulted in zeolite with larger-pore size (Okuhara, 2002) but with mesopores in which can significantly minimize the diffusion problem
weaker acidic strength (Chung et al., 2008). Thus, although zeolite for reactants to access into the active sites of the catalyst (Mbaraka
with larger-pore size may overcome the diffusion limitation problem and Shanks, 2006). Apart from that, the physical and chemical
but the reaction rate is still rather slow due to low acidity strength. properties of these mesoporous materials can be manipulated by
Brito et al. (2007) reported the application of different zeolite Y incorporating suitable organic or inorganic functional groups into the
catalysts in transesterification of waste cooking oil. Several types of mesoporous silica matrix (Mbaraka et al., 2006). For instance, in order
zeolite Y with different concentration of Al2O3 and Na2O were utilized to obtain a solid acid catalyst, organosulfonic groups can be
in the reaction, however, the catalysts were found to give poor incorporated onto the mesoporous silica material as shown in
performance. The highest biodiesel yield attained was only 26.6% Fig. 13 (Melero et al.). The organosulfonic acid anchored on
although the reaction was carried out at high reaction temperature of mesoporous silica acts as Brønsted acid (active sites) that is suitable
460 °C, methanol to oil molar ratio of 6 and reaction time of 22 min. to catalyze esterification and transesterification reactions. Mesos-
Similar reports on the low activity of zeolite catalyst when used in tructure silica materials functionalized with propylsulfonic acid
transesterification were also reported in recent publications (Shu et (SO3H) groups have been reported to have good catalytic activity in
al., 2007, Kiss et al., 2006, Okuhara, 2002). esterification of refined and unrefined oil (Melero et al.; Melero et al.,
2009).
6.5.5. Sulfonic ion-exchange resin To date, application of sulfonated silica in transesterification of
Ion-exchange resins are insoluble macroporous polymer that is waste cooking oil is still limited. Most of the recent studies reported
capable to exchange specific ions within the polymer itself with other focused on transesterification of refined oil rather than oil with high
ions in a solution or reaction media. Normally, sulfonic ion-exchange FFA and waste cooking oil. Mbaraka et al. (2006) reported the
resins are co-polymers of divinylbenzene (DVB), styrene and sulfonic utilization of acidic mesoporous silica in esterification of beef tallow.
acid groups (as the active sites-Brønsted acidity) (Özbay et al., 2008). From the study, it was found that propylsulfonic acid-functionalized
The polymer structure of the resin is mainly characterized by the mesoporous silica (SBA-15-SO3H) has an excellent catalytic activity in
composition of the cross-linking component (normally DVB), which esterification. Nearly 95% conversion of FFA from beef tallow was
will then determine its surface area and pore size distribution attained at relatively short reaction time of 30 min, at reaction
(Pääkkönen and Krause, 2003). Besides that, their catalytic activity temperature of 120 °C and methanol to FFA ratio of 20. Unfortunately,
is also strongly dependent on their swelling properties because the the catalytic activity of SBA-15-SO3H reduced drastically for the
swelling capacity limits reactant accessibility to the acid sites and thus subsequent reaction cycles due to accumulation of organic or
affects their overall activity (Feng et al., 2010). Common types of carbonaceous matter on the catalyst surface that blocks the acidic
acidic ion-exchange resin are such as Amberlyst-15, Amberlyst-35 sites. However, leaching of the active sites was rule out as the reason
and Nafion SAC-13. These catalysts were reported to give good for catalyst deactivation since the authors overcome this problem by
performance in FFA esterification, but, weak in transesterification incorporating hydrophobic agent (propyltrimethoxysilane) onto SBA-
(Chen et al., 1999; Vicente et al., 1998). 15-SO3H in order to increase its hydrophobicity. This bi-functionized
Application of Amberlyst-15 with acidic functional groups has (acidic and hydrophobic) mesostructure silica would have lower
exhibited excellent catalytic activity in esterification reaction. Kiss et pores polarity, thereby reducing the interaction between polar
al. tested the activity of several solid acid catalysts in the esterification impurities (contained in the beef tallow) with the sulfonic acid
of dodecanoic acid with 2-ethylhexanol at 150 °C. Amberlyst-15 was groups. Results showed that high FFA conversion (84%) was
512 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Fig. 13. Organosulfonic acid-functionalized mesoporous silica.

maintained for the second cycle of reaction. However, in-depth producing acid-based catalyst. However, it should be noted that the
research and analysis work using sulfonic modified silica in carbon-based catalyst could not be prepared by sulfonation of an
transesterification reaction (typically from high free fatty acid oil) incomplete carbonized resin, amorphous glassy carbon, activated carbon
should be carried out to justify the potential of this solid acid catalyst. or natural graphite (Hara, 2009). Heating these carbon materials with
H2SO4 will only result to low density SO3H groups and do not function as
6.5.7. Sulfonated carbon-based catalyst a solid acid catalyst.
Carbon-based catalyst is classified as a type of sugar catalyst, in which Up to date, there are only minimum study that reported on the use
sugar, starch or cellulose is carbonized incompletely at temperature of carbon-based material for biodiesel production, typically from
below 500 °C (Hara, 2009). The incomplete carbonized carbon is then waste cooking oil. In a recent study reported by Lou et al., various
immersed in concentrated sulfuric acid, H2SO4 (purity N96%) and heated sulfonated carbon-based catalysts were derived, mainly from starch,
to 150 °C for 15 h under the flow of N2 (Lou et al., 2008, Takagaki et al., cellulose, sucrose and D-glucose (Lou et al., 2008). It was found that
2006, Nakajima et al., 2007). The resulting catalyst was claimed to be the carbon-based catalyst derived from starch has a relatively larger-pore
first of its kind and is called as sulfonated carbon-based catalyst which volume (0.81 cm3/g) and pore size (8.2 nm) allowing better reactants
comprise high density functional group such as sulfonic group (SO3H) access to the SO3H sites. Apart from that, starch derived carbon-based
and carboxyl group (COOH) (Hara, 2009). The structure of a typical catalyst also have a relatively higher total acid sites (1.97 mmol/g)
carbon-based catalyst derived from D-glucose is schematically shown in and higher sulfur content (5.9 wt.%). In addition, high FAME yield
Fig. 14 (Takagaki et al., 2006). The minimum unit in this material is a (92%) was attained from transesterification of waste cooking oil at
nanographene sheet (ca. 1 nm) comprised of 10–20 carbon six- reaction temperature of 80 °C, methanol to oil molar ratio of 30,
membered rings (Hara, 2009). In addition, the sulfonated carbon catalyst loading of 10 wt.% (referred to weight of oil) and reaction
material has a high Hammett acid strength (H0) of −8 to −11, which time of 8 h. Furthermore, the catalyst was found to be stable even after
is almost comparable to concentrated H2SO4 (Okamura et al., 2006). fifty cycles of repeated reaction. Nevertheless, this catalyst must still
Furthermore, leaching of SO3H group was not observed in the be further improved such as proper optimization work on the catalyst
esterification of high free fatty acid oil (Takagaki et al., 2006). Such preparation as well as transesterification reaction conditions.
finding not only open up a new route to minimize extensively the usage
of H2SO4 in the industries, but also lead to a greener approach for 6.5.8. Heteropolyacids (HPAs)
Heteropolyacids (HPAs) catalyst has attracted researcher's atten-
tion recently due to their excellent water tolerant ability, poses strong
Brønsted acidity (stronger than conventional homogeneous acid,
H2SO4) and high catalytic activity and stability (Sivasamy et al., 2009,
Narasimharao et al., 2007). Typical HPAs which are easily available are
H3PW12O40, H4SiW12O40, H3PMo12O40 and H4SiMo12O40 (Zhang et al.,
2010). In addition, adding appropriate ratio of salt (Cs+, NH+ 4 and
Ag+) to HPAs will dramatically increase its surface area and allow
easier accessibility of reactant to its active sites (Narasimharao et
al., 2007). However, it should be noted that HPAs can be slightly
soluble in the reaction media and resulting to homogeneous
reaction, in which contributed to the overall reaction rate (Sivasamy
et al., 2009). Therefore, leaching of active sites may also easily occur
and cause serious catalyst deactivation.
Application of HPAs in biodiesel synthesis from waste cooking oil
was reported by Cao et al. (2008). H3PW12O40·6H2O (PW12) was used
as the HPAs catalyst and the waste cooking oil contained high amount
of FFA (15.65%) and water content (0.1%). Optimum yield of 87%%
biodiesel was attained at reaction temperature of 65 °C, methanol to
oil molar ratio of 70:1 and 14 h of reaction time. At the same time, 4 Å
zeolite was introduced into the reaction media as an adsorbent to
adsorb water. Although PW12 has a high tolerance towards FFA
content and is stable even after 5 reaction cycles, however, relatively
Fig. 14. Proposed schematic structure of the carbon materials prepared from D-glucose.
high methanol to oil molar ratio and long reaction time may restrict
The materials are amorphous carbons consisting of polycyclic aromatic carbon sheets the application of this catalyst in industrial scale. Apart from that, the
with phenolic OH and COOH groups in addition to SO3H groups. catalyst was found not stable when reaction temperature was
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 513

increased more than 60 °C. The authors claimed that this observation given for this observation. On the other hand, in a solvent-free
is due to the nature of waste cooking oil that contains many system (without the addition of hexane), biodiesel yield dropped
undesirable compounds that will cause side reactions when carrying significantly to merely 20–65% when methanol or ethanol were used
out transesterification reaction at higher temperature. as primary alcohol. Nevertheless, if branched primary alcohol
On the other hand, attempt was made to synthesize HPA with (isopropanol) was used in a solvent-free system, high yield of
higher acidity by introducing Lewis acid into HPA, making it contains biodiesel (97%) could be maintained (Nelson et al., 1996). The main
both Brønsted and Lewis acid sites. This was done by loading Brønsted advantage of using secondary alcohol in biodiesel production is to
acidic HPAs on Lewis acidic supporters, such as ZrO2, TiO2 or Ta2O5 reduce the solidification point and subsequently improve its cloud
(Zhang et al., 2009). Proper coordination of a Lewis acid to a Brønsted point and pour point characteristic (Salis et al., 2005).
acid could enhance its original acidity and therefore forming a bi-
functional acid site catalysts. Zhang et al. (2009) studied the potential 6.6.2. Pseudomonas cepacia (PS 30)
of this double acidic sites of HPAs in transesterification of waste Other than Lipozym IM 60, Nelson et al. (1996) also screened the
cooking oil. Zr0.7H0.2PW12O40 (ZrHPW) with nanotube structure was potential of PS 30 in transesterification of tallow to biodiesel.
successfully synthesized using natural cellulose fiber as a template. However, the efficiency of the enzyme in catalyzing the transester-
ZrHPW exhibited high acidity capacity with 1350 μmol/g acid sites ification reaction was low, even when solvent was introduced into the
was detected, much higher than its original HPW (892 μmol/g). This reaction mixture. Biodiesel yield obtained was merely 13.9–28.8% for
high density acid sites is comparable to recent studies reported in the primary alcohols (methanol, ethanol and isobutanol) and 44.1% for
literature, such as carbonized glucose (1550 μmol/g) and mesoporous secondary alcohol (isopropanol). The reaction conditions reported
sulfated silica zirconia (1260 μmol/g). Apart from that, high biodiesel are; reaction temperature at 45 °C, stirring speed of 200 rpm, 5 h
yield of 98.9% was attained at reaction temperature of 65 °C, methanol reaction time, 0.34 molar of triglyceride in hexane, methanol to oil
to oil molar ratio of 20, catalyst loading of 2.1% and reaction time of molar ratio of 3 and 12.5–25% enzyme by weight of tallow.
8 h. Moreover, after 5 reaction cycles, biodiesel yield still remain at Although Nelson et al. reported that PS 30 was not a good enzyme,
95%. nevertheless, Wu et al. still attempted to study the same enzyme but
by optimizing the transesterification reaction using response surface
6.6. Enzyme (biocatalyst) catalyzed transesterification methodology (RSM) (Wu et al., 1999). In his study, restaurant grease
was used as feedstock and ethanol with 95% purity was used as the
Enzymatic transesterification especially those using lipase has source of primary alcohol. Using the regression equation developed, it
drawn researcher's attention in last ten years due to the downstream was predicted that an optimum biodiesel yield of 85.4% can be
processing problem posed by chemical transesterification. Huge obtained at the following optimum conditions: 38.4 °C, 2.47 h, 13.7%
amount of wastewater generation and difficulty in glycerol recovery lipase (PS-30), and grease to ethanol molar ratio of 1:6.6. However,
are among problems that eventually increase the overall biodiesel the apparent yield of biodiesel obtained was way below 85% when
production cost and being not environmental benign. In contrast, using the predicted optimum conditions. Consequently, the authors
enzyme catalysis proceeds without the generation of by-products, improved the biodiesel yield by introducing a second portion of lipase
easy recovery of product, mild reaction condition, insensitive to high PS-30 into the reaction mixture after 1 h of reaction. Unfortunately,
FFA oil and catalyst can be reuse (Kulkarni and Dalai, 2006). These this method did not improve the biodiesel yield significantly. On the
advantages prove that enzyme catalyzed biodiesel production has other hand, if the second portion was changed to lipase SP 435
high potential to be an eco-friendly process and a promising (Candida antarctica) with 5% (wt) of loading, 96% of ethyl ester yield
alternative to the chemical process. However, it still has its fair can be obtained. Nevertheless, neither lipase PS-30 nor SP435 when
share of constraints especially when implemented in industrial scale used individually could give high yield as predicted using RSM.
such as high cost of enzyme, slow reaction rate and enzyme Recently, enzymatic transesterification reached another important
deactivation (Bajaj et al., 2010). The following section reviews some milestone with the use of immobilization technology. The purpose of
of the potential enzymes which have been studied in transesterifica- immobilization is to provide a more rigid external backbone for lipase
tion of waste cooking oil to biodiesel with emphasis given on the molecule which will result in a faster reaction rate (Knezevic et al.,
reaction and optimum condition used. 1998). Lipase can be immobilized into ion-exchange resin, photo-
cross linkable resin, silica beads and alumina through adsorption,
6.6.1. Mucor miehei (Lipozym IM 60) cross-linking, entrapment and covalent bonding method. The advan-
In a study by Nelson et al., Lipozym IM 60 was found to show a tages of immobilized enzymes over free enzymes are easier lipase
promising result in transesterification of tallow with high FFA recovery, high stability, insensitive to solvent, and reusable. Hsu et al.
content. Based on the study, two different types of alcohols were immobilized PS-30 within a phyllosilicate sol–gel matrix (IM PS-30)
investigated: (1) primary alcohols, such as methanol, ethanol, and was subsequently subjected to transesterification using waste
propanol, butanol and isobutanol and (2) secondary alcohol, such grease oil (Hsu et al., 2002). It was found that IM PS-30 was able to
as isopropanol and 2-butanol. The effect of adding solvent (hexane) convert 84–94% of grease oil to biodiesel for both primary and
into the reaction mixture was also studied. The purpose of secondary alcohols. The reaction was conducted at 50 °C, alcohol to
introducing solvent into enzymatic transesterification process is to grease molar ratio of 1:4, 100 mg IM PS-30, and reaction time of 18 h.
increase the solubility between methanol and glycerol and therefore In addition, the enzymatic activity of IM PS-30 did not deactivate even
minimizes the possibility of enzyme deactivation caused by after 48 h of reaction as the lipase was strongly constrained within the
methanol and glycerol (Halim and Harun Kamaruddin, 2008). It matrix. Furthermore, by adding molecular sieve (to remove water) in
was found that Lipozym IM 60 is a promising enzyme in the reaction mixture, the biodiesel yield increased by as much as 20%
transesterification of tallow with primary alcohol resulting in with shorter reaction time.
biodiesel yield as high as 93–99%. The optimum reaction conditions
were: reaction temperature at 45 °C, stirring speed of 200 rpm, 5 h 6.6.3. C. antarctica (Novozym 435)
reaction time, 0.34 molar of triglyceride in hexane, methanol to oil C. antarctica was first used by Nelson et al. (1996) in transester-
molar ratio of 3 and 12.5–25% enzyme (by weight of tallow). ification of high FFA tallow to produce biodiesel. It was observed that
However, for secondary alcohol, the yield of biodiesel obtained was Novozym 435 has high enzymatic activity when secondary alcohol
relatively low. Only 19–24% of biodiesel yield was obtained using the (2-butanol) was used in a solvent-free system. Biodiesel yield of
same reaction condition as primary alcohols. No explanation was 96.4% was obtained at the following reaction conditions: 0.34 molar
514 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

of tallow, reaction temperature at 45 °C, alcohol to tallow molar ratio 2007). B. subtilis was initially encapsulated within the net of
of 3:1, stirring speed of 200 rpm and reaction time of 16 h. Besides hydrophobic carrier with magnetic particles (Fe3O4), and then the
that, adding suitable amount of water (6 mol% based on tallow) into secreted lipase can be conjugated with carboxyl at the magnetic poly-
the system was found to promote ester formation when secondary microsphere surface. This magnetic cell biocatalyst (MCB) was
alcohol was used with Novozym 435 (Nelson et al., 1996). This is claimed to have better dispersion during transesterification and can
because water is an essential element to maintain the catalytic easily be separated from the reaction mixture by subjecting to an
activity of enzymes. However, the presence of too much water may external magnetic field. From the study, it was found that biodiesel
cause hydrolysis of oil in which is undesirable and eventually will yield can reached up to 90% at reaction temperature 40 °C, pH 6.5,
lead to lower conversion of oil to biodiesel. loading of 3.0% MCB, adding methanol in two stepwise and 72 h
In another study reported by Watanabe et al. (2001), stepwise reaction time. Furthermore, MCB can be easily regenerated without
methanol addition was introduced during transesterification in order losing its enzymatic activity.
to avoid deactivation of enzyme by methanol and to prolong its
durability. Waste cooking oil was used as the main feedstock whereas
immobilized Novozym 435 was used as the enzyme. The reaction was 6.6.5. Rhizopus oryzae
conducted with three different ways: (1) three-step batch (fixed flow Chen et al. investigated the enzymatic conversion of waste cooking
rate) methanolysis, (2) three-step flow methanolysis and (3) one- oil using immobilized R. oryzae lipase (Chen et al., 2006). A three-step
step flow methanolysis. For the three-step batch reaction, the amount batch transesterification reactor was used and stepwise process was
of methanol used was divided equally in each step. 90.4% conversion introduced in the reactor to reduce the poisoning of enzyme by
of waste oil to biodiesel was obtained after the third step. In addition, methanol. The optimum reaction condition was reported at 40 °C,
for the three-step continuous feed, methanol at a flow rate of 6.1, 6.3 methanol to oil molar ratio of 4, immobilized lipase to oils weight ratio
and 4.2 mL/h was introduced into the first, second and third reactor, of 30%, pressure of 1 atm and reaction time of 30 h. Biodiesel yield in
respectively and 90.9% of biodiesel conversion was achieved. On the the range of 88–90% can be obtained under these conditions.
other hand, for one-step methanolysis, waste oil was diluted initially
with 90% methyl ester (biodiesel) with weight ratio of 3:1 and
equimolar amount of methanol, relative to total fatty acids in the 6.6.6. Penicillium expansum
waste oil. The mixture was then fed into a bioreactor containing 3 g of Immobilized P. expansum on resin D4020 was reported by Li et al.
Novozym 135 at a flow rate of 4 mL/h and 90% of biodiesel conversion (2009) to have high enzymatic activity in transesterification of waste
was achieved. oil to biodiesel. The process was further enhanced with the addition of
Apart from that, in order to reduce the overall production cost of adsorbents such as molecular sieve and blue silica gel into the reaction
enzyme, Novozym sp. 99–135 was immobilized on low-cost textile mixture with the purpose to absorb excess water produced during
cloth (Chen et al., 2009). The immobilization carrier of textile cloth esterification of FFA with methanol. Excess water will not only result
was initially activated by mixing with co-fixing agents consisted of to aggregation of the enzyme in hydrophobic media, hence reducing
polyethylene glycol, tween and span (surfactants), gelatin and its enzymatic activity, but also have a negative effect on the enzyme's
lecithin. The lipase solution was then blended with activated stability. Optimum biodiesel yield of 92.8%% was obtained when the
immobilization carrier and dried in air prior to use. Three reactors reaction was carried out at 35 °C, stirring speed of 200 rpm, 2 g of
are connected in series to form a three-step reaction system in waste oil, 0.4 g t-amyl alcohol, 0.96 g blue silica gel, 168 U immobi-
which glycerol was separated at each step. Waste cooking oil was lized penicillium expansum and 7 h of reaction time. One molar
used as the main feedstock with high acid value of 143 mg KOH/g. equivalent of methanol was added at reaction time of 0, 1 and 3 h.
The optimum reaction condition of lipase/hexane/water/waste Apart from that, the synthesized enzyme displayed higher stability in
cooking oil weight ratio of 25:15:10:100, reaction temperature of waste oil with 68.4% of the original enzymatic activity retained even
45 °C and reactant flow of 1.2 ml/min gave 91% yield of biodiesel. after recycle and reuse for 10 batches.
However, for this continuous process, biodiesel yield was found to
decrease to 76% after running the reaction at optimum condition
continuously for 100 h. The expected reasons are: (1) the glycerol 7. Other methods or technologies for biodiesel production
absorbed on the surface of immobilized lipase restricts mass
transfer between substrate and enzyme, and (2) enzyme was Currently, most commercial scale biodiesel plants use batch or
poisoned by methanol. continuous-type reactor to produce biodiesel from refined vegetable
Most researchers prefer to use hexane as co-solvent in enzymatic oils. In addition, homogeneous base catalysts such as potassium
transesterification as it can enhance higher activity and posed good hydroxide and sodium hydroxide are the most widely used catalysts
stability. However, hexane is a hydrophobic solvent in which due to their fast reaction rate and mild reaction conditions. However,
hydrophilic compounds used as subtract (alcohol) or obtained by- if waste cooking oil is to be used as the main feedstock to produce
product (glycerol) are easily immiscible in hydrophobic reaction biodiesel, a two- step process may be required in which the first step
medium. This has resulted to low solubility of both mediums and high is esterification process to reduce FFA content in the oil and the
possibility to deactivate the enzyme (Halim and Harun Kamaruddin, second step is transesterification process. However, this two-step
2008). In a study done by Halim and Harun Kamaruddin, tert-butanol process generates a lot of waste water and is not environmental
was used as an ideal solvent in a reaction mixture containing waste benign. Alternatively, as presented earlier, heterogeneous and
cooking palm oil, methanol and Novozym 435. High solubility of enzymatic catalysts have the potential to overcome the problems
methanol and glycerol in tert-butanol can eventually eliminate the posed by homogeneous catalysts. However, most of the studies
negative effect of methanol and glycerol on enzyme activity. From the reported are carried out at laboratory scale. Therefore, if heteroge-
study, it was found that 88% of biodiesel yield can be obtained at neous or enzymatic based transesterification process need to be
reaction temperature 40 °C, methanol to oil molar ratio of 4:1, 4% scaled up to industrial scale, mass and heat transfer limitation must be
Novozym and 12 h reaction time. carefully addressed. The following section reviews some of the
potential technologies that can facilitate heterogeneous and enzy-
6.6.4. Bacillus subtilis matic transesterification system to overcome the mass and heat
Ying et al. was the first group of researchers that used B. subtilis for transfer limitation and thus obtaining higher yield of biodiesel in a
transesterification of waste cooking oil to biodiesel (Ying and Chen, shorter reaction time.
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 515

7.1. Oscillatory flow reactor (OFR) for transesterification reaction scale to industrial production scale and process safety (Vyas et al.,
2010).
Oscillatory flow reactor (OFR) was first introduced by Harvey et al. A few reports concerning the usage of microwave irradiation in
(2003) to produce biodiesel through some improvement in mixing transesterification have been reported recently (Lertsathapornsuk et
intensity between reactants. OFR is a novel type of continues flow al., 2008, Barnard et al., 2007, Leadbeater and Stencel, 2006). One of
reactor, consisting of tubes containing equally spaced orifice plate the interesting studies is the development of a continuous flow
baffles. Therefore, an oscillatory motion is superimposed upon the net microwave reactor to produce biodiesel from waste cooking oil as
flow of the process fluid, creating flow patterns conducive for efficient reported by Barnard et al. (2007). In that study, transesterification
heat and mass transfer, whilst maintaining plug flow regime (Harvey reaction was performed using a commercially available multimode
et al., 2003). In addition, each baffle essentially behaves as a stirred microwave apparatus (CEM MARS). Initially, a 10 L mixture of waste
tank that lead to excellent mixing and suspension by creating vortices cooking oil, methanol and catalyst was prepared (1:6 ratio of oil to
between orifice baffles and oscillating fluid (Zheng et al., 2007). This is methanol and 1 wt.% KOH as catalyst) and placed in a holding tank.
an essential element in designing a biodiesel reactor especially when The mixture was then pumped into the microwave reactor vessel at a
heterogeneous catalysts are used due to the presence of three flow rate of 2 L/min and heated to 50 °C using microwave power of
immiscible phases (oil–alcohol–catalyst) at the initial stage of 1600 W. After 10 min, the products were pumped out from the
reaction. Thus, improvement in mixing and suspension of catalysts reactor. It was found that the overall biodiesel conversion achieved
tend to produce higher yield of biodiesel in a shorter reaction time 97.9%. Furthermore, when the flow rate of the reaction mixture was
compared to conventional batch-type stirred tank reactor. Apart from increased to 7.2 L/min, 98.9% of biodiesel conversion can still be
that, OFR allows longer residence time as the mixing is independent of obtained. Apart from that, the study also reported the overall energy
the net flow and hence the reactor length-to-diameter ratio can be consumption by microwave irradiation and conventional heating as
reduced. This is an important plus point if the process is scaled up for shown in Table 7. The results clearly show that microwave irradiation
commercial application in order to reduce the overall capital and process would be significantly more energy-efficient than conven-
pumping cost. tional heating in a continuous biodiesel production process.
Harvey et al. applied OFR in the production of biodiesel from waste
cooking oil and pure rapeseed oil (Harvey et al., 2003). The reaction 7.3. Ultrasonic technology in transesterification reaction
was performed at temperature of 20–70 °C, residence time of 10–
30 min and molar ratio of methanol to oil was maintained at 1.5. Pure Ultrasonic technology has been recognized as an effective method
sodium hydroxide (32.4 g) was dissolved in pure methanol initially at to enhance mass transfer rate between immiscible liquid–liquid
40 °C for 1 h. It was found that at 50 °C and 30 min of reaction time, phases within a heterogeneous system (Ji et al., 2006). Therefore, it
nearly 99% of biodiesel was produced. Moreover, the product contains has been widely used in various biological and chemical reactions to
negligible amount of triglyceride and diglyceride. However, some improve the yield within a shorter reaction time. Ultrasound is
traces of monoglyceride were detected (Noureddini and Zhu, 1997). defined as sound with frequency beyond human ear can respond. The
Nevertheless, Harvey et al. (2003) concluded that in-depth study on normal sound frequency that can be detected by human lies between
OFR in transesterification with heterogeneous catalyst is promptly 16 and 18 kHz, but frequency for ultrasound generally lies between
required as OFR is ideal for suspending solid catalysts or polymer 20 kHz and 100 MHz (Vyas et al., 2010). This high frequency sound
supported catalysts. wave will compresses and stretches the molecular spacing of a
medium in which it passes through. Thus, molecules will be
7.2. Microwave technology in transesterification reaction continuously vibrated and cavities will be created. As a result, micro
fine bubbles are formed through sudden expansion and collapse
In recent years, electromagnetic energy utilization and develop- violently, generating energy for chemical and mechanical effects
ment has gained much interest by many research groups (Corsaro et (Colucci et al., 2005). Furthermore, the collapsed bubbles will disrupt
al., 2004). Microwave irradiation is one of the examples as this process the phase boundary and impinging of the liquids to create micro jets,
posed several advantages such as higher yields of cleaner product, leading to intensively emulsification of the system (Ji et al., 2006).
minimum energy consumption and environmentally benign com- Ultrasonic technology in transesterification has proven to be an
pared to conventional heating in various chemical reactions (Grois- efficient mixing tool and provides sufficient activation energy to
man and Gedanken, 2008). In fact, conventional heating process initiate the reaction (Singh et al., 2007). Ultrasonic-assisted transes-
suffers significant drawback due to its limitation such as it is terification does not only shorten reaction time, but also minimize the
dependent on the thermal conductivity of materials, specific heat molar ratio of alcohol to oil and reduce energy consumption
and density (Groisman and Gedanken, 2008). Apart from that,
conventional heating is rather slow and heat is not distributed
Table 7
uniformly in a reaction vessel, resulting to more energy (than the
Energy estimation for the preparation of biodiesel using conventional and microwave
theoretical value) is required for a particular chemical reaction heating.
(Mutyala et al., 2010). Moreover, direct contact between the hot
Entry Reaction conditions Energy consumption (kJ/L)a
reaction vessel surface with reaction media (reactants) may results to
product decomposition especially when heated for long period of 1 Conventional heatingb 94.3
time. In contrast, microwaves transfer energy in a form of electro- 2 Microwave continuous flow 26.0
(7.2 L/min feedstock flow)
magnetic and not thermal heat reflux. The oscillating microwave field
3 Microwave continuous flow 60.3 (92.3)d
tends to oscillate polar ends of molecules or ions continuously (Marra (2 L/min feedstock flow)c
et al., 2010; Azcan and Danisman, 2007). Consequently, collisions and 4e Microwave heating (4.6 L batch reaction) 90.1
friction between the moving molecules is created and generate heat a
Normalized for energy consumed per liter of biodiesel prepared.
(Marra et al., 2010). Heat is therefore directly deposited into the b
On the basis of values from the joint U.S. Department of Agriculture and U.S.
reaction media and resulted to rapid temperature increase through- Department of Energy 1998 study into the life cycle inventory of biodiesel and
out the sample (Liu and Cheng, 2009, Azcan and Danisman, 2007). petroleum diesel for use in an urban bus.
c
Assuming a power consumption of 1700 W and a microwave input of 1045 W.
Thus, higher yield of products can be obtained in a short reaction time. d
Assuming a power consumption of 2600 W and a microwave input of 1600 W.
However, the major drawbacks of using microwave irradiation for e
Assuming a power consumption of 1300 W, a microwave input of 800 W, a time to
biodiesel synthesis are the scaling-up of the process from laboratory reach 50 °C of 3.5 min, and a hold time at 50 °C of 1 min.
516 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

compared to conventional mechanical stirring method (Vyas et al., with a mechanical agitator and circulation pump that is capable to
2010). However, up to now, there are only few studies applying homogenize the mixture sufficiently even without the use of co-
ultrasonic technology in transesterification of waste cooking oil solvent. Furthermore, it is more economical to improve the mass
(Refaat and El Sheltawy, 2008, Wang et al., 2007). Wang et al. transfer limitation through mechanical stirring rather than using co-
explored the potential of ultrasonic technology in enzymatic solvent as additional energy is required to distillate the co-solvent.
transesterification of high acid value waste oil (Wang et al., 2007). Nevertheless, more studies are required to verify the effect of co-
Commercial immobilized Novozym 435 from C. antarctica was utilized solvent in heterogeneous transesterification system particularly for
as a biocatalyst in the system. It was found that the enzymatic activity enzyme catalyst.
was enhanced with the assistance of low frequency and mild energy
ultrasonic sound wave. Under the optimal ultrasonic assistant 8. Conclusion
reaction conditions, such as 8% oil quantity of Novozym 435, molar
ratio of propanol to oil 3:1, frequency of ultrasonic assistant 28 kHz Biodiesel is a renewable and alternative fuel to petroleum-based
occupied at power 100 W, reaction temperature at 40–45 °C, an diesel that is non-toxic, biodegradable and does not contribute net
overall biodiesel (propyl oleate) conversion of 94.86% was achieved in carbon emission to the atmosphere. Currently, biodiesel is produced
only 50 min. Furthermore, ultrasonic sound wave tends to reduce the through transesterification reaction from vegetable oils such as
adsorption of biodiesel and glycerol on the surface of immobilized rapeseed, soybean and palm oil. However, the high prices of these
Novozym 435. As a result, Novozym 435 can be recycle to use with oils in the global market have sharply increased the overall biodiesel
clean appearances, well decentralizations, no agglomeration, easy production cost and making it not economically viable as compared to
washing and well operational stability. petrol based diesel. Furthermore, the oils are important commodities
in the human food supply chain and therefore its conversion to
7.4. Co-solvent biodiesel in a long run may not be sustainable. As an alternative, this
paper has addressed the potential of using waste cooking oil, as a
Transesterification reaction is a slow reaction process which cheap and economical feedstock for biodiesel production.
typical requires 30 min to few hours to drive the reaction towards Nevertheless, current commercial technology (homogeneous base
completion, depending on the type of catalyst used. This is because catalyst) was found not suitable for the transesterification of waste
the reactants used in transesterification (triglyceride and alcohol) are cooking oil due to its high FFA content. On the other hand, using
not miscible and therefore causing mass transfer limitation especially homogeneous acid catalyst requires longer reaction time and could
at the initial time of the reaction. In order to accelerate the rate of potentially cause corrosion on equipment. Therefore, it is clear that
transesterification reaction, researchers have recently introduced co- research on heterogeneous catalyst, either base or acid type, should be
solvent such as tetrahydrofuran (THF), hexane and diethyl ether carried out extensively to develop a suitable catalyst to convert waste
(DME) into the reaction mixture with the aim to increase the cooking oil to biodiesel with special emphasis on catalyst deactivation
solubility and subsequently improve the mass transfer rate between and regeneration. Apart from that, enzymatic transesterification is
oil and methanol phase (Guan et al., 2009). Experimental results have another possible way to produce biodiesel from waste cooking oil due
shown that THF is indeed a good co-solvent that can accelerate to its high stability towards FFA content in oils. However, the enzyme
biodiesel production within a shorter reaction time, either for should be synthesized in a cheaper way and available for commercial
homogeneous or heterogeneous system (Chai et al., 2007, Yang and use.
Xie, 2007, Peña et al., 2009). However, one of the disadvantages of Apart from the development of suitable heterogeneous and
using THF is the co-solvent must be separated out from the final enzymatic catalysts for biodiesel production from waste cooking oil,
product upon completion of reaction. Although it can be easily one major drawback still exist. Up-scaling of laboratory scale process
distilled out together with methanol, however separation between involving immiscible phases to commercial scale is not an easy task
methanol and THF becomes rather problematic since both chemicals due to mass and heat transfer limitation. Nevertheless, recent
have close boiling point. Note that distilled methanol should be advances in technologies such as oscillatory flow reactor (OFR),
further purified before it is subjected to the next cycle of reaction. In microwave irradiation, ultrasonic technology and co-solvent have
another study, biodiesel itself was used as co-solvent since it is shown high potential in overcoming the limitation. These technolo-
miscible in oil and methanol phase (Park et al., 2009). Furthermore, gies not only facilitate transesterification reaction in terms of
biodiesel does not need to be separated out since it itself is the final increasing mixing intensity, heat and mass transfer rate, but also
product in the transesterification reaction. In the study by Park et al., it proved to be more energy-efficient as compared to conventional
was reported that when biodiesel was added as co-solvent, transes- heating process. Thus, in order to materialize the technology for
terification reaction can be completed within a relatively shorter converting waste cooking oil to biodiesel using heterogeneous and
reaction time (20 min) as compared to system without co-solvent enzyme catalysts, extensive research work on process scaling-up and
(60 min). more advance research on developing sustainable catalysts should be
Nevertheless, when co-solvent was used in heterogeneous conducted simultaneously.
transesterification of waste cooking oil, a contra observation was
reported (Guan et al., 2009). Result showed that the addition of co- Acknowledgements
solvent THF or DME into the reaction media, caused a drop in biodiesel
yield. Based on this finding, the authors suggested that mass transfer The authors would like to acknowledge the funding given by
between the methanol and oil phase in a heterogeneous catalytic Universiti Sains Malaysia (Research University Grant No. 1001/
system was not affected significantly even with the addition of co- PJKIMIA/814062, Short-term Grant No. 304/PJKIMIA/6039015, Re-
solvent. Furthermore, the authors observed that when THF was added search University Postgraduate Research Grant Scheme No. 1001/
into the reaction mixture, the solid catalyst easily stick with glycerol PJKIMIA/8031018 and USM Fellowship) for this project.
and agglomerate on the reactor wall. Thus, catalyst was deactivated as
a result of agglomeration. A similar observation was also reported in a References
homogeneous system when waste cooking oil was utilized as the
feedstock (Sabudak and Yildiz, 2010). The authors concluded that the Azcan N, Danisman A. Alkali catalyzed transesterification of cottonseed oil by
microwave irradiation. Fuel 2007;86:2639–44.
positive effect of THF addition was found to be insignificant because Bajaj A, Lohan P, Jha PN, Mehrotra R. Biodiesel production through lipase catalyzed
the reactor used for the transesterification reaction was equipped transesterification: an overview. J Mol Catal B Enzym 2010;62:9-14.
M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518 517

Balaji BS, Chanda BM. Simple and high yielding syntheses of β-keto esters catalysed by Guan G, Kusakabe K, Yamasaki S. Tri-potassium phosphate as a solid catalyst for
zeolites. Tetrahedron 1998;54:13237–52. biodiesel production from waste cooking oil. Fuel Process Technol 2009;90:520–4.
Barnard TM, Leadbeater NE, Boucher MB, Stencel LM, Wilhite BA. Continuous-flow Gui MM, Lee KT, Bhatia S. of edible oil vs. non-edible oil vs. waste edible oil as biodiesel
preparation of biodiesel using microwave heating. Energy Fuels 2007;21:1777–81. feedstock. Energy 2008;33:1646–53.
Brito A, Borges ME, Otero N. Zeolite Y as a heterogeneous catalyst in biodiesel fuel Gutierrez-Baez R, Toledo-Antonio JA, Cortes-Jacome MA, Sebastian PJ, Vazquez A.
production from used vegetable oil. Energy Fuels 2007;21:3280–3. Effects of the SO4 groups on the textural properties and local order deformation of
Canakci M, Van Gerpen J. Biodiesel production via acid catalysis. Trans ASAE (Am Soc SnO2 rutile structure. Langmuir 2004;20:4265–71.
Agric Eng) 1999;42:1203–10. Haas MJ, McAloon AJ, Yee WC, Foglia TA. A process model to estimate biodiesel
Canakci M, Van Gerpen J. A pilot plant to produce biodiesel from high free fatty acid production costs. Bioresour Technol 2006;97:671–8.
feedstocks. Trans ASAE (Am Soc Agric Eng) 2003;46:945–54. Halim SFA, Harun Kamaruddin A. Catalytic studies of lipase on FAME production from
Cantrell DG, Gillie LJ, Lee AF, Wilson K. Structure–reactivity correlations in MgAl waste cooking palm oil in a tert-butanol system. Process Biochem 2008;43:1436–9.
hydrotalcite catalysts for biodiesel synthesis. Appl Catal A 2005;287:183–90. Hara M. Environmentally benign production of biodiesel using heterogeneous catalysts.
Cao F, Chen Y, Zhai F, Li J, Wang J, Wang X. Biodiesel production from high acid value ChemSusChem 2009;2:129–35.
waste frying oil catalyzed by superacid heteropolyacid. Biotechnol Bioeng Harvey AP, Mackley MR, Seliger T. Process intensification of biodiesel production using
2008;101:93-100. a continuous oscillatory flow reactor. J Chem Technol Biotechnol 2003;78:338–41.
Chai F, Cao F, Zhai F, Chen Y, Wang X, Su Z. Transesterification of vegetable oil to Heidekum A, Harmer MA, Hoelderich WF. Addition of carboxylic acids to cyclic olefins
biodiesel using a heteropolyacid solid catalyst. Adv Synth Catal 2007;349:1057–65. catalyzed by strong acidic ion-exchange resins. J Catal 1999;181:217–22.
Chen X, Xu Z, Okuhara T. Liquid phase esterification of acrylic acid with 1-butanol Hsu AF, Jones K, Foglia TA, Marmer WN. Immobilized lipase-catalysed production of
catalyzed by solid acid catalysts. Appl Catal A 1999;180:261–9. alkyl esters of restaurant grease as biodiesel. Biotechnol Appl Biochem 2002;36:
Chen G, Ying M, Li W. Enzymatic conversion of waste cooking oils into alternative fuel — 181–6.
biodiesel. Appl Biochem Biotechnol 2006;132:911–21. International Energy Agency I. Key world energy statistic; 2006.
Chen H, Peng B, Wang D, Wang J. Biodiesel production by the transesterification of International Energy Agency I. Key world energy statistic; 2008.
cottonseed oil by solid acid catalysts. Front Chem Eng Chin 2007;1:11–5. Jacobson K, Gopinath R, Meher LC, Dalai AK. Solid acid catalyzed biodiesel production
Chen Y, Xiao B, Chang J, Fu Y, Lv P, Wang X. Synthesis of biodiesel from waste cooking oil from waste cooking oil. Appl Catal B 2008;85:86–91.
using immobilized lipase in fixed bed reactor. Energy Convers Manage 2009;50: Ji J, Wang J, Li Y, Yu Y, Xu Z. Preparation of biodiesel with the help of ultrasonic and
668–73. hydrodynamic cavitation. Ultrasonics 2006;44:411–4.
Chin LH, Hameed BH, Ahmad AL. Process optimization for biodiesel production from Jitputti J, Kitiyanan B, Rangsunvigit P, Bunyakiat K, Attanatho L, Jenvanitpanjakul P.
waste cooking palm oil (Elaeis guineensis) using response surface methodology. Transesterification of crude palm kernel oil and crude coconut oil by different solid
Energy Fuels 2009;23:1040–4. catalysts. Chem Eng J 2006;116:61–6.
Chung KH, Chang DR, Park BG. Removal of free fatty acid in waste frying oil by Kapias T, Griffiths RF. Spill behaviour using REACTPOOL — part I. Results for accidental
esterification with methanol on zeolite catalysts. Bioresour Technol 2008;99: releases of chlorosulphonic acid (HSO3Cl). J Hazard Mater 2001;81:19–30.
7438–43. Khder AS, El-Sharkawy EA, El-Hakam SA, Ahmed AI. Surface characterization and
Colucci JA, Borrero EE, Alape F. Biodiesel from an alkaline transesterification reaction of catalytic activity of sulfated tin oxide catalyst. Catal Commun 2008;9:769–77.
soybean oil using ultrasonic mixing. J Am Oil Chem Soc 2005;82:525–30. Kiss AA, Dimian AC, Rothenberg G. Solid acid catalysts for biodiesel production —
Corma A, Garcia H. Organic reactions catalyzed over solid acids. Catal Today 1997;38: towards sustainable energy. Adv Synth Catal 2006;348:75–81.
257–308. Knezevic ZD, Siler-Marinkovic SS, Mojovic LV. Kinetics of lipase-catalyzed hydrolysis of
Corsaro A, Chiacchio U, Pistarà V, Romeo G. Microwave-assisted chemistry of palm oil in lecithin/izooctane reversed micelles. Appl Microbiol Biotechnol
carbohydrates. Curr Org Chem 2004;8:511–38. 1998;49:267–71.
Cvengros J, Cvengrosova Z. Used frying oils and fats and their utilization in the Kouzu M, Kasuno T, Tajika M, Sugimoto Y, Yamanaka S, Hidaka J. Calcium oxide as a
production of methyl esters of higher fatty acids. Biomass Bioenergy 2004;27: solid base catalyst for transesterification of soybean oil and its application to
173–81. biodiesel production. Fuel 2008a;87:2798–806.
de Almeida RM, Noda LK, Gonçalves NS, Meneghetti SMP, Meneghetti MR. Kouzu M, Kasuno T, Tajika M, Yamanaka S, Hidaka J. Active phase of calcium oxide used
Transesterification reaction of vegetable oils, using superacid sulfated TiO2-base as solid base catalyst for transesterification of soybean oil with refluxing methanol.
catalysts. Appl Catal A 2008;347:100–5. Appl Catal A 2008b;334:357–65.
Demirbas A. Biodiesel from waste cooking oil via base-catalytic and supercritical Kouzu M, Yamanaka SY, Hidaka JS, Tsunomori M. Heterogeneous catalysis of calcium
methanol transesterification. Energy Convers Manage 2009;50:923–7. oxide used for transesterification of soybean oil with refluxing methanol. Appl Catal
Demirbas MF, Balat M. Recent advances on the production and utilization trends of bio- A 2009;355:94–9.
fuels: a global perspective. Energy Convers Manage 2006;47:2371–81. Kulkarni MG, Dalai AK. Waste cooking oil — an economical source for biodiesel: a
Di Serio M, Ledda M, Cozzolino M, Minutillo G, Tesser R, Santacesaria E. Transester- review. Ind Eng Chem Res 2006;45:2901–13.
ification of soybean oil to biodiesel by using heterogeneous basic catalysts. Ind Eng Kumar Tiwari A, Kumar A, Raheman H. Biodiesel production from jatropha oil (Jatropha
Chem Res 2006;45:3009–14. curcas) with high free fatty acids: an optimized process. Biomass Bioenergy
Dos Reis SCM, Lachter ER, Nascimento RSV, Rodrigues Jr JA, Reid MG. Transesterification 2007;31:569–75.
of Brazilian vegetable oils with methanol over ion-exchange resins. J Am Oil Chem Lam MK, Lee KT, Mohamed AR. Sulfated tin oxide as solid superacid catalyst for
Soc 2005;82:661–5. transesterification of waste cooking oil: an optimization study. Appl Catal B
Escobar JC, Lora ES, Venturini OJ, Yáñez EE, Castillo EF, Almazan O. Biofuels: 2009a;93:134–9.
environment, technology and food security. Renewable Sustainable Energy Rev Lam MK, Tan KT, Lee KT, Mohamed AR. Malaysian palm oil: surviving the food versus
2009;13:1275–87. fuel dispute for a sustainable future. Renewable Sustainable Energy Rev 2009b;13:
Exxon Mobile. A report on energy trends, green house gas emissions and alternative 1456–64.
energy; 2004. Leadbeater NE, Stencel LM. Fast, easy preparation of biodiesel using microwave heating.
Faria EA, Marques JS, Dias IM, Andrade RDA, Suarez PAZ, Prado AGS. Nanosized and Energy and Fuels 2006;20:2281–3.
reusable SiO2/ZrO2 catalyst for highly efficient biodiesel production by soybean Lertsathapornsuk V, Pairintra R, Aryusuk K, Krisnangkura K. Microwave assisted in
transesterification. J Braz Chem Soc 2009;20:1732–7. continuous biodiesel production from waste frying palm oil and its performance in
Felizardo P, Neiva Correia MJ, Raposo I, Mendes JF, Berkemeier R, Bordado JM. a 100 kW diesel generator. Fuel Process Technol 2008;89:1330–6.
Production of biodiesel from waste frying oils. Waste Manage 2006;26:487–94. Leung DYC, Guo Y. Transesterification of neat and used frying oil: optimization for
Feng Y, He B, Cao Y, Li J, Liu M, Yan F, et al. Biodiesel production using cation-exchange biodiesel production. Fuel Process Technol 2006;87:883–90.
resin as heterogeneous catalyst. Bioresour Technol 2010;101:1518–21. Li NW, Zong MH, Wu H. Highly efficient transformation of waste oil to biodiesel by
Freedman B, Pryde EH, Mounts TL. Variables affecting the yields of fatty esters from immobilized lipase from Penicillium expansum. Process Biochem 2009;44:
transesterified vegetable oils. J Am Oil Chem Soc 1984;61:1638–43. 685–8.
Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterification of oils. J Liu CZ, Cheng XY. Microwave-assisted acid pretreatment for enhancing biogas
Biosci Bioeng 2001;92:405–16. production from herbal-extraction process residue. Energy Fuels 2009;23:6152–5.
Furuta S, Matsuhashi H, Arata K. Biodiesel fuel production with solid superacid catalysis Lotero E, Liu Y, Lopez DE, Suwannakarn K, Bruce DA, Goodwin Jr JG. Synthesis of
in fixed bed reactor under atmospheric pressure. Catal Commun 2004a;5:721–3. biodiesel via acid catalysis. Ind Eng Chem Res 2005;44:5353–63.
Furuta S, Matsuhashi H, Arata K. Catalytic action of sulfated tin oxide for etherification Lou WY, Zong MH, Duan ZQ. Efficient production of biodiesel from high free fatty acid-
and esterification in comparison with sulfated zirconia. Appl Catal A 2004b;269: containing waste oils using various carbohydrate-derived solid acid catalysts.
187–91. Bioresour Technol 2008;99:8752–8.
Gazmuri AM, Bouchon P. Analysis of wheat gluten and starch matrices during deep-fat Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol 1999;70:1-15.
frying. Food Chem 2009;115:999-1005. Marra F, De Bonis MV, Ruocco G. Combined microwaves and convection heating: a
Gerhard Knothe JVG, Krahl Jurgen. The biodiesel handbook. AOCS Press; 2004. conjugate approach. J. Food Eng 2010;97:31–9.
Granados ML, Poves MDZ, Alonso DM, Mariscal R, Galisteo FC, Moreno-Tost R, et al. Matsuhashi H, Miyazaki H, Kawamura Y, Nakamura H, Arata K. Preparation of a solid
Biodiesel from sunflower oil by using activated calcium oxide. Appl Catal B superacid of sulfated tin oxide with acidity higher than that of sulfated zirconia and
2007;73:317–26. its applications to aldol condensation and benzoylation. Chem Mater 2001;13:
Groisman Y, Gedanken A. Continuous flow, circulating microwave system and its 3038–42.
application in nanoparticle fabrication and biodiesel synthesis. J J Phys Chem C Mbaraka IK, Shanks BH. Conversion of oils and fats using advanced mesoporous
2008;112:8802–8. heterogeneous catalysts. J Am Oil Chem Soc 2006;83:79–91.
Gryglewicz S. Rapeseed oil methyl esters preparation using heterogeneous catalysts. Mbaraka IK, McGuire KJ, Shanks BH. Acidic mesoporous silica for the catalytic
Bioresour Technol 1999;70:249–53. conversion of fatty acids in beef tallow. Ind Eng Chem Res 2006;45:3022–8.
518 M.K. Lam et al. / Biotechnology Advances 28 (2010) 500–518

Melero J.A., Bautista L.F., Morales G., Iglesias J., Sánchez-Vázquez R. Biodiesel production Singh AK, Fernando SD, Hernandez R. Base-catalyzed fast transesterification of soybean
from crude palm oil using sulfonic acid-modified mesostructured catalysts. Chem oil using ultrasonication. Energy Fuels 2007;21:1161–4.
Eng J. doi:10.1016/j.cej.2009.12.037. Sivasamy A, Cheah KY, Fornasiero P, Kemausuor F, Zinoviev S, Miertus S. Catalytic
Melero JA, Iglesias J, Morales G. Heterogeneous acid catalysts for biodiesel production: applications in the production of biodiesel from vegetable oils. ChemSusChem
current status and future challenges. Green Chem 2009;11:1285–308. 2009;2:278–300.
Miao CX, Gao Z. Preparation and properties of ultrafine SO2− 4 /ZrO2 superacid catalysts. Suarez PAZ, Plentz Meneghetti SM, Meneghetti MR, Wolf CR. Transformation of
Mater Chem Phys 1997;50:15–9. triglycerides into fuels, polymers and chemicals: some applications of catalysis in
Mittelbach M, Enzelsberger H. Transesterification of heated rapeseed oil for extending oleochemistry. Quim Nova 2007;30:667–76.
diesel fuel. J Am Oil Chem Soc 1999;76:545–50. Takagaki A, Toda M, Okamura M, Kondo JN, Hayashi S, Domen K. Esterification of higher
MPOC. World's oils and fats production share in 2007. Malaysian Palm Oil Council fatty acids by a novel strong solid acid. Catal Today 2006;116:157–61.
(MPOC); 2008. TBW. Biofuels: the promise and risks. The World Bank; 2008.
Muniyappa PR, Brammer SC, Noureddini H. Improved conversion of plant oils and Tesser R, Di Serio M, Guida M, Nastasi M, Santacesaria E. Kinetics of oleic acid
animal fats into biodiesel and co-product. Bioresour Technol 1996;56:19–24. esterification with methanol in the presence of triglycerides. Ind Eng Chem Res
Mutyala S, Fairbridge C, Paré JRJ, Bélanger JMR, Ng S, Hawkins R. Microwave 2005;44:7978–82.
applications to oil sands and petroleum: a review. Fuel Process Technol 2010;91: Thurmond W. BIODIESEL 2020: Global market survey, feedstock trends and market
127–35. forecasts. Emerging Markets Online 2008. Available at: http://www.emerging-
Nag A. Biofuels refining and performance. McGraw Hill; 2008. markets.com/biodiesel/default.asp (accessed April 2009).
Nakajima K, Hara M, Hayashi S. Environmentally benign production of chemicals and Toledo-Antonio JA, Gutierrez-Baez R, Sebastian PJ, Vazquez A. Thermal stability and
energy using a carbon-based strong solid acid. J Am Ceram Soc 2007;90:3725–34. structural deformation of rutile SnO2 nanoparticles. J Solid State Chem 2003;174:
Narasimharao K, Brown DR, Lee AF, Newman AD, Siril PF, Tavener SJ. Structure–activity 241–8.
relations in Cs-doped heteropolyacid catalysts for biodiesel production. J Catal Vasudevan PT, Briggs M. Biodiesel production — current state of the art and challenges. J
2007;248:226–34. Ind Microbiol Biotechnol 2008;35:421–30.
Nawar WW. Chemical changes in lipids produced by thermal processing. J Chem Educ Velasco J, Dobarganes C. Oxidative stability of virgin olive oil. Eur J Lipid Sci Technol
1984;61:299–302. 2002;104:661–76.
Nelson LA, Foglia TA, Marmer WN. Lipase-catalyzed production of biodiesel. J Am Oil Vicente G, Coteron A, Martinez M, Aracil J. Application of the factorial design of
Chem Soc 1996;73:1191–5. experiments and response surface methodology to optimize biodiesel production.
Noureddini H, Zhu D. Kinetics of transesterification of soybean oil. J Am Oil Chem Soc Ind Crops Prod 1998;8:29–35.
1997;74:1457–63. Vyas AP, Verma JL, Subrahmanyam N. A review on FAME production processes. Fuel. 89,
Okamura M, Takagaki A, Toda M, Kondo JN, Domen K, Tatsumi T. Acid-catalyzed 1–9.
reactions on flexible polycyclic aromatic carbon in amorphous carbon. Chem Mater Wan Omar WNN, Nordin N, Mohamed M, Amin NAS. A two-step biodiesel production
2006;18:3039–45. from waste cooking oil: optimization of pre-treatment step. J Appl Sci 2009;9:
Oku T, Nonoguchi M, Moriguchi T, Oku T, Nonoguchi M, Moriguchi T. Method of 3098–103.
producing of fatty alkyl esters and/or glycerine and fatty acid alkyl ester-containing Wang Y, Ou S, Liu P, Xue F, Tang S. Comparison of two different processes to synthesize
composition. PCT Application No WO2005/021697; 2005. biodiesel by waste cooking oil. J Mol Catal A: Chem 2006;252:107–12.
Okuhara T. Water-tolerant solid acid catalysts. Chem Rev 2002;102:3641–66. Wang JX, Huang QD, Huang FH, Wang JW, Huang QJ. Lipase-catalyzed production of
Omota F, Dimian AC, Bliek A. Fatty acid esterification by reactive distillation: part 2 — biodiesel from high acid value waste oil using ultrasonic assistant. Chin J Biotechnol
kinetics-based design for sulphated zirconia catalysts. Chem Eng Sci 2003;58: 2007;23:1121–8.
3175–85. Watanabe Y, Shimada Y, Sugihara A, Tominaga Y. Enzymatic conversion of waste edible
Özbay N, Oktar N, Tapan NA. Esterification of free fatty acids in waste cooking oils oil to biodiesel fuel in a fixed-bed bioreactor. J Am Chem Soc 2001;78:703–7.
(WCO): role of ion-exchange resins. Fuel 2008;87:1789–98. Wu WH, Foglia TA, Marmer WN, Phillips JG. Optimizing production of ethyl esters of
Pääkkönen PK, Krause AOI. Diffusion and chemical reaction in isoamylene etherification grease using 95% ethanol by response surface methodology. J Am Oil Chem Soc
within a cation-exchange resin. Appl Catal A 2003;245:289–301. 1999;76:517–21.
Park YM, Lee DW, Kim DK, Lee JS, Lee KY. The heterogeneous catalyst system for the Xavier NM, Lucas SD, Rauter AP. Zeolites as efficient catalysts for key transformations in
continuous conversion of free fatty acids in used vegetable oils for the production of carbohydrate chemistry. J Mol Catal A: Chem 2009;305:84–9.
biodiesel. Catal Today 2008;131:238–43. Xie W, Peng H, Chen L. Calcined Mg–Al hydrotalcites as solid base catalysts for
Park JY, Kim DK, Wang ZM, Lee JS. Fast biodiesel production with one-phase reaction. methanolysis of soybean oil. J Mol Catal A: Chem 2006;246:24–32.
Appl Biochem Biotechnol 2009;154:246–52. Yadav GD, Murkute AD. Preparation of a novel catalyst UDCaT-5: enhancement in
Park YM, Lee JY, Chung SH, Park IS, Lee SY, Kim DK, et al. Esterification of used vegetable activity of acid-treated zirconia — effect of treatment with chlorosulfonic acid
oils using the heterogeneous WO3/ZrO2 catalyst for production of biodiesel. vis-à-vis sulfuric acid. J Catal 2004;224:218–23.
Bioresour Technol 2010;101:S59–61. Yan S, Salley SO, Simon Ng KY. Simultaneous transesterification and esterification of
Patil P, Deng S, Isaac Rhodes J, Lammers PJ. Conversion of waste cooking oil to biodiesel unrefined or waste oils over ZnO–La2O3 catalysts. Appl Catal A 2009;353:203–12.
using ferric sulfate and supercritical methanol processes. Fuel 2010;89:360–4. Yang Z, Xie W. Soybean oil transesterification over zinc oxide modified with alkali earth
Peña R, Romero R, Martínez SL, Ramos MJ, Martínez A, Natividad R. Transesterification metals. Fuel Process Technol 2007;88:631–8.
of castor oil: effect of catalyst and co-solvent. Ind Eng Chem Res 2009;48:1186–9. Ying M, Chen G. Study on the production of biodiesel by magnetic cell biocatalyst based
Peng BX, Shu Q, Wang JF, Wang GR, Wang DZ, Han MH. Biodiesel production from waste on lipase-producing Bacillus subtilis. Appl Biochem Biotechnol 2007;137–140:
oil feedstocks by solid acid catalysis. Process Saf Environ Prot 2008;86:441–7. 793–803.
Pimentel D, Pimentel M. Global environmental resources versus world population Zabeti M, Wan Daud WMA, Aroua MK. Activity of solid catalysts for biodiesel
growth. Ecol Econ 2006;59:195–8. production: a review. Fuel Process Technol 2009;90:770–7.
Ramadhas AS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA rubber Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste cooking oil:
seed oil. Fuel 2005;84:335–40. 1. Process design and technological assessment. Bioresour Technol 2003a;89:1-16.
Refaat AA, El Sheltawy ST. Comparing three options for biodiesel production from waste Zhang Y, DubeÌ MA, McLean DD, Kates M. Biodiesel production from waste cooking oil: 2.
vegetable oil. WIT Trans Ecol Environ 2008:133–40. Economic assessment and sensitivity analysis. Bioresour Technol 2003b;90:229–40.
Sabudak T, Yildiz M. Biodiesel production from waste frying oils and its quality control. Zhang X, Li J, Chen Y, Wang J, Feng L, Wang X. Heteropolyacid nanoreactor with double
Waste Manage. doi:10.1016/j.wasman.2010.01.007. acid sites as a highly efficient and reusable catalyst for the transesterification of
Sahoo PK, Das LM, Babu MKG, Naik SN. Biodiesel development from high acid value waste cooking oil. Energy Fuels 2009;23:4640–6.
polanga seed oil and performance evaluation in a CI engine. Fuel 2007;86:448–54. Zhang S, Zu YG, Fu YJ, Luo M, Zhang DY, Efferth T. Rapid microwave-assisted
Salis A, Pinna M, Monduzzi M, Solinas V. Biodiesel production from triolein and short transesterification of yellow horn oil to biodiesel using a heteropolyacid solid
chain alcohols through biocatalysis. J Biotechnol 2005;119:291–9. catalyst. Bioresour Technol 2010;101:931–6.
Sharma YC, Singh B. Development of biodiesel: current scenario. Renewable Zheng S, Kates M, Dubé MA, McLean DD. Acid-catalyzed production of biodiesel from
Sustainable Energy Rev 2009;13:1646–51. waste frying oil. Biomass Bioenergy 2006;30:267–72.
Shu Q, Yang B, Yuan H, Qing S, Zhu G. Synthesis of biodiesel from soybean oil and Zheng M, Skelton RL, Mackley MR. Biodiesel reaction screening using oscillatory flow
methanol catalyzed by zeolite beta modified with La3+. Catal Commun 2007;8: meso reactors. Process Saf Environ Prot 2007;85:365–71.
2159–65. Zhou A, Thomson E. The development of biofuels in Asia. Appl Energy 2009;86:11–20.

You might also like