Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

DOI: 10.1002/cssc.

200800253

Catalytic Applications in the Production of Biodiesel from Vegetable Oils


Arumugam Sivasamy,[a, b] Kien Yoo Cheah,[c] Paolo Fornasiero,[a, d] Francis Kemausuor,[a, e] Sergey Zinoviev,[a] and Stanislav Miertus*[a]

278

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 278 300

The predicted shortage of fossil fuels and related environmental concerns have recently attracted significant attention to scientific and technological issues concerning the conversion of biomass into fuels. First-generation biodiesel, obtained from vegetable oils and animal fats by transesterification, relies on commercial technology and rich scientific background, though continuous progress in this field offers opportunities for improvement. This Review focuses on new catalytic systems for

the transesterification of oils to the corresponding ethyl/ methyl esters of fatty acids. It also addresses some innovative/ emerging technologies for the production of biodiesel, such as the catalytic hydrocracking of vegetable oils to hydrocarbons. The special role of the catalyst as a key to efficient technology is outlined, together with the other important factors that affect the yield and quality of the product, including feedstock-related properties and various system conditions.

1. Introduction
Notwithstanding the recent debate on the perspectives of the further development of first-generation biofuels (based on energy crops) due to alarming socio-economic implications (feedstock competition with food), and even environmental concerns in some cases (non-sustainable practices leading to the increase of CO2 levels), the production of biodiesel from vegetable oils and animal fats still attracts significant and justified attention. Considering the existing agricultural and industrial production capacities, as well as the spinning interest in the exploitation of future-generation feedstocks such as nonedible, waste, and algae oils, it is expected that a significant part of future renewable fuels will still be represented by vegetable oil/animal fat based biodiesel. A variety of economic and social benefits for local communities, especially in developing countries, have been put forward to support the wider adoption of biodiesel.[15] The production costs of diesel fuel from petroleum, which reached 1.00 USD L1 (crude oil and refining) in May 2008,[6] have been becoming evermore competitive with those of biodiesel. Very different costs for biodiesel production have been reported, ranging from 0.42 USD L1 (biodiesel from animal fats, New Zealand; 2007) to 0.90 USD L1 (oilseed rape, Europe; soybean, USA; palm oil, Malaysia; 2007),[7] or even 2.00 USD L1 for next-generation biodiesel (algae, NREL estimate).[5] As a rule, 7680 %[8] or a larger fraction[9] of the biodiesel from vegetable oils cost accounts for the feedstock cost. For example, the production costs of 1.75 USD L1 for biodiesel produced from rape oil include the price of rape oil at 1.60 USD L1.[10] On the other hand, in the case of waste oil the situation can be inverse, for example, 0.53 USD L1 bio-feedstock cost versus 0.64 USD L1 processing costs.[11] The feedstock cost can also vary largely from one bio-feedstock to another, for example, 1.28 USD L1 for rapeseed oil versus 0.70 USD L1 for soybean oil[5] versus 0.13 USD L1 for yellow grease.[12] In addition, the bio-feedstock cost varies substantially with time (e.g. the price of palm oil rose from 0.40 USD L1 in 2004 to over 1.00 USD L1 in 2008) and also regionally (e.g. the costs of biodiesel production from rapeseed oil are 0.55 USD L1 in Korea against 2.85 USD L1 in Japan).[5] Even though the portion related to the processing and conversion of the feedstock in the total product cost is often negligible, the role of the production technology is not to be underestimated. In fact, specific technological approaches are required to treat multiple feedstocks, including low-cost ones like waste oils. In addition, in the proChemSusChem 2009, 2, 278 300

spective of further biodiesel market development there is a tendency of equilibration of feedstock prices which will bring about a tight competition between production know-how. Apart from the significant role of sound practices in feedstock exploitation and cultivation in meeting global environmental goals and in stabilizing the socio-economic balance of the biodiesel industry, the further development of the biodiesel industry also lies in scientific and technological innovation.[13, 14] In this regard, there is a significant interest in improving the existing biodiesel production methods from both economic and environmental viewpoints, as well as in alternative and innovative production processes. The direct use of straight vegetable oils (SVOs) as fuels in diesel engines were investigated well before the energy crisis of the 1970s.[15] However, the high viscosity of straight vegetable oils[16, 17] causes a range of problems and leads to poor fuel characteristics. Among possible solutions, such as engine modifications, blending straight vegetable oils with fossil diesel, micro-emulsification, transesterification, or thermal cracking, the transesterification of oils to the corresponding fatty acid methyl esters (FAMEs) has seen wider application owing to its easy, cost-effective technology as well as because of the favorable physicochemical parameters of the fuel.[13, 18] Another promising approach to convert vegetable oils and fats into quality fuel is hydrotreating/cracking technology, which
[a] Dr. A. Sivasamy, Prof. P. Fornasiero, F. Kemausuor, Dr. S. Zinoviev, Prof. S. Miertus Area of Pure and Applied Chemistry International Centre for Science and High Technology United Nations Industrial Development Organization (ICS-UNIDO) Area Science Park, Padriciano 99, 34012 Trieste (Italy) Fax: (+ 39) 040-9228115 E-mail: stanislav.miertus@ics.trieste.it [b] Dr. A. Sivasamy Chemical Engineering Department Central Leather Research Institute Adyar, Chennai-600 02 (India) [c] Dr. K. Y. Cheah Malaysian Palm Oil Board (MPOB) Persiaran Institusi Bandar Baru Bangi, 43000 Kajang, Selangor (Malaysia) [d] Prof. P. Fornasiero Chemistry Department and ICCOM-CNR University of Trieste Via L. Giorgieri 1, 34127 Trieste (Italy) [e] F. Kemausuor Department of Agricultural Engineering Kwame Nkrumah University of Science and Technology KNUST Post Office, Kumasi (Ghana)

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

279

S. Miertus et al. produces the diesel fuel with a chemical composition similar to that of conventional diesel (hydrocarbons). Both types of biodiesel can replace petroleum diesel without the need for modification or adjustment of engines. In addition reducing CO2 emissions, biodiesel offers significant environmental benefits such as the lack of SOx emissions as a result of the virtual absence of sulfur in the feedstock. In the case of FAME biodiesel, the higher oxygen content with respect to petroleum diesel ensures more complete combustion and a lower content of CO, hydrocarbons, and particulate in emissions.[1923] In addition, FAME biodiesel is non-toxic and biodegradable, though its biodegradability results in lower stability of the fuel. Sometimes, its oxidation should be prevented by adding additives, which is particularly the case of biodiesel obtained from vegetable oils which have been previously subjected to extraction procedures to recover high-value components, such as vitamins, which have antioxidant properties.[24, 25]The continuing growth in demand of biodiesel following several legislations in a number of countries that have mandated biodiesel blends in transportation fuels has been a challenge for scientists to develop innovative and more efficient processes for producing biodiesel from vegetable oils. In this context, the development of new catalytic systems for the transesterification and hydrocracking of oils remains a challenging research frontier.

Arumugam Sivasamy received his MPhil (1988) and PhD (1992) from the University of Madras (India). He then worked at the Indian Institute of Toxicological Research (Lucknow, 1992 1996) before joining the Central Leather Research Institute (Chennai) as a senior scientist (1996present). He also worked as a fellow (20062007) and researcher (2008) at the ICS-UNIDO. His research interests include biomass valorization and utilization and environmental catalysis. Kien-Yoo Cheah completed his BEng at the University of Malaya (Malaysia, 1982) and received his PhD from the University of Cambridge (UK, 1996). Since 1984, he has been a researcher at the Malaysian Palm Oil Board (MPOB) and is currently Head of Milling and Processing Unit. His main research focus is biodiesel. His other research interests include catalysis, process development in palm-oil refining, processing, and oleochemicals.

Francis Kemausuor holds an MPhil in Engineering for Sustainable Development from the University of Cambridge (UK). He is a lecturer at the Department of Agricultural Engineering at Kwame Nkrumah University of Science and Technology (KNUST; Ghana) and a research fellow at The Energy Centre at the same university. He was a fellow at the ICS-UNIDO during 2008. His research interests lie in technology transfer for bioenergy systems.

Sergey Zinoviev obtained his PhD from the N. D. Zelinsky Institute of Organic Chemistry of the Russian Academy of Sciences in 2002. His research interests are focused on catalytic applications in green chemistry and environmental technologies. He was appointed Associate Scientific Officer of the Area of Chemistry at the ICSUNIDO in 2007. He is currently the acting principal investigator of the newly established laboratory of catalysis for the production of biofuels and biobased chemicals at the ICS-UNIDO. Stanislav Miertus graduated from the Slovak Technological University (STU; Bratislava, Slovakia) in 1971. He then worked as a researcher at the Polymer Institute of the Academy of Sciences (19711981), completing his PhD there in 1975, before joining the Faculty of Chemistry at STU (19811990). He has been Area Chief of ICS-UNIDO since 1998 and has coordinated programs on catalysis and sustainable chemistry, biofuels and biobased products, and environmentally degradable plastics, among other topics.

Paolo Fornasiero completed his PhD in 1996. After one year as postdoctoral fellow at the University of Reading (UK, 1997), he joined the University of Trieste (Italy) as Assistant Professor in 1998. In 2006, he was appointed Associate Professor of Inorganic Chemistry at the same university. He has been the scientific director of the CNRICCOM Trieste research unit since 2008. His scientific interests are focused on technological applications of materials science and environmental heterogeneous catalysis.

280

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 278 300

Catalytic Applications in Biodiesel Production

2. Transesterification Reaction
Transesterification refers to the reaction of an ester group with an alcohol that has a structure different to that of the original alcohol moiety of the ester, such that a new ester group is formed in which the original alcohol moiety is exchanged with that of the reacting alcohol. In the case of the transesterification of triglycerides of fatty acids (vegetable oils) with methanol (classical biodiesel production process), the three ester groups of a triglyceride molecule in which three fatty acid moieties are attached to a single alcohol moiety (i.e. that of glycerol) react with three molecules of methanol to yield three molecules of esters each containing single fatty acid and methanol moieties and one molecule of glycerol (Scheme 1). Therefore, the general chemical name of biodiesel produced by the transesterification of vegetable oils is fatty acid methyl esters (i.e. FAME biodiesel).

Scheme 1. Transesterification reaction of a triglyceride with methanol.

Methanol is commonly used in industrial biodiesel production as a result of its relatively low cost and easy availability. Ethanol is not commonly used owing to a high cost related to the removal the final 4 % water. As the non-catalytic transesterification is too slow and energetically unfavorable, acid or base catalysts are used. For the reaction mechanisms of the transesterification of triglycerides with acid and base catalysts, see Scheme 2 and Scheme 3, respectively. Various catalysts can be used, such as alkaline hydroxides and methoxides, inorganic acids and their salts, transition-metal compounds, silicates, and lipases. The classical reaction protocol for the transesterification of triglycerides with methanol using homogeneous catalysts such as sodium methoxide requires mixing and stirring the reagents in a batch reactor. At the end of the reaction, the non-polar phase containing the ester and the polar phase containing glycerol and methanol are separated to recover the products, catalyst, and the excess of methanol. After the ester is separated, other purification steps may be required, for example, washing, distillation, and/or extraction to remove remaining glycerol and other impurities. The excess of methanol recovered from the glycerol phase is usually reused, and the byproduct glycerol can also be valorized to improve the economics of the process.

Scheme 2. General reaction mechanism of transesterification of triglycerides with acid catalysts.

2.1. Kinetics and Thermodynamics of Triglyceride Transesterification While the general features of the reaction mechanism of acidand base-catalyzed transesterification are well known, not many studies have been directed towards better understanding the principles and parameters governing the kinetics and thermodynamics of the transesterification of vegetable oils, which is instrumental in improving the commercial performance of the process. The transesterification of vegetable oil is a complex process; the reaction rate and equilibrium yields are affected by numerous chemical and physical factors. Vegetable oil is composed from over 100 substances, and different oils have different compositions that can vary even for the same oil. In addition, the most common industrial transesterification of triglycerides with methanol is normally a heterogeneous process, which involves at least two immiscible liquid phases (oil and methanol phases) or even one solid phase if a solid catalyst is used. Mass-transfer processes over liquidliquid and liquidsolid interfaces often complicate the understanding of the true reactivity by kinetic analysis. In fact, several authors indicated that the transesterification of vegetable oil with methanol represents a two-phase system in the first few minutes of

ChemSusChem 2009, 2, 278 300

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

281

S. Miertus et al. diglyceride methanol monoglyceride ester K 2 k2 =k2 2 monoglyceride methanol glycerol ester K 3 k3 =k3 3 Most studies show that the equilibrium constant (K3) for the third step of the conversion of the monoglyceride into the corresponding ester is one to two orders of magnitude higher than those of the preceding steps. For comparison, the values of rate and equilibrium constants for the three different reaction steps are reported in Table 1. It is also a common tendency that there is an increase in the value of the forward rate constant from step 1 to step 3. In most cases, second-order kinetics has been observed at 6:1 methanol/oil molar ratio (double excess of alcohol per reacting ester group in the triglyceride). Depending on the systems, second-order,[26] pseudo-second-order,[30, 31] or, at higher excess of MeOH, pseudo-first-order[18] kinetic models have been successfully applied for data fitting. Several mechanistic models have been proposed, all of which resulted in quite suitable fitting of the experimental data by the respective mathematical models; namely, a simple sequence of reversible transesterification steps,[18] consecutive steps with contribution of the shunt mechanism (simultaneous reaction of all three ester groups in the triglyceride),[18, 26] and competitive hydrolysis transesterification.[28] In most cases, the authors claim that the observed reaction rates and calculated constants are indeed reflecting the chemical processes and not mass-transfer processes. For example, activation energies in the range of 5 20 kcal mol1 have been observed for separate transesterification steps.[18, 26, 30] In some studies, induction periods have been observed on the kinetic curves which were attributed to the mass-transfer limitations in the initial reaction phase.[26, 30] In particular, mixing[26] has been outlined as critical for the reaction to proceed in the kinetic regime from the beginning. At optimal mixing intensity (600 rpm), the lag, which was observed on kinetic curves at lower stirring rates, disappeared. The authors[26] indicated that the mass-transfer contribution to the kinetic behavior is felt at the beginning of reaction because of the phase-transfer limitations. As the methyl ester is accumulated, it acts as a mutual solvent and ultimately a single-phase system is formed which is favorable for a kinetically controlled

Scheme 3. General reaction mechanism of transesterification of triglycerides with base catalysts.

the reaction that then becomes a single phase as a sufficient amount of methyl esters is accumulated which act as a mutual solvent;[26] however, it was also noted that after some glycerol is formed, the phase-separation phenomenon takes place again.[27] Side reactions, such as hydrolysis, also often take place. All these phenomena make the equilibrium and rate constants as well as the general kinetic behavior very dependent on the particular system. Generally, for either an acid- or base-catalyzed mechanism, the transesterification of a triglyceride is a reversible process, where the overall process is composed of a sequence of three equilibrium steps involving the formation of intermediate diand monoglycerides, and finally of a glycerol molecule [see Eq. (1)(3)]. Each component process is an equilibrium described by its equilibrium constant Ki, forward rate constant ki, and reverse rate constant ki. triglyceride methanol diglyceride ester K 1 k1 =k1 1

Table 1. A comparison of rate and equilibrium constants of component processes in the transesterification of triglycerides from several reported studies. Vegetable oil Rapeseed Pongamia Palm Palm Soybean Soybean k1 k2 k3 k1 k2 k3 K1[a] [L mol1 min1] [L mol1 min1] [L mol1 min1] [L mol1 min1] [L mol1 min1] [L mol1 min1] 5.01 0.029 0.036[b] 0.011 0.050 3.822[d] 4.93 0.0058 0.070[b] 0.018 0.215 1.215[d] 29.7 0.011 0.141[b] 0.131 0.242 792[d] 3.55 0.014 0.000 0.110 2.99 0.021 0.082 1.228 0.79 0.00051 0.002 0.007 1.06 1.99 K2[a] K3[a] T [8C] KOH Ref. [%] 23 60 60 60 50 60 1.5 12 1 1[c] [28] [29] [30] [31] [26] [18]

1.71 132.8 0.27 21.74

[a] Calculation of equilibrium constants based on the final concentrations. [b] Units: (wt % min)1. [c] NaOH. [d] Units: min1.

282

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 278 300

Catalytic Applications in Biodiesel Production process. A low reaction rate after the induction phase was also suggested[32] to be connected to the two-phase reaction and a low oil content in the methanol phase at the beginning of the reaction. In another study,[30] it was shown that a lag period on the kinetic curves could be a result of a low catalyst concentration. The lag on the kinetic curves disappeared at KOH concentrations of 1 mol % and higher. A model describing the kinetics of the base-catalyzed transesterification of rapeseed oil was proposed,[28] which included the competition of the transesterification reaction with the irreversible side process of saponification of methyl esters and glycerides with KOH. A good (78 % probability) approximation of experimental data by the model was reached. Some antagonistic effects have been noted, namely the larger contribution of the saponification process at higher base concentrations and higher temperatures. The influence of water was also shown to be critical in the reaction of cotton seed oil with ethanol,[33] whereby a higher water content resulted in shifting the equilibrium of all the three steps to the left (towards the reactants). For example, the conversion of triglyceride was less than 50 % at a water content of 4.4 % in ethanol and an ethanol-to-oil ratio of 4:1. It was also noted that quite a high ratio of ethanol to oil (9:1) was required to achieve good conversions even at low water content. The authors, however, did not provide conclusions on the nature of the inhibition effect of water in ethanol. In the case of enzymatic transesterification, the kinetics are usually described by the MichaelisMenten model,[34] which is typical for most single substrate enzymes. For lipase-catalyzed esterification, transesterification, and interesterification processes, the ping-pong bi-bi model with competitive substrate inhibition is most often applied.[3537] As inhibition of the enzyme activity by various system components is a typical feature in enzyme-catalyzed reactions, specific models have been developed to describe the kinetic data and, therefore, the reaction conditions have to be adjusted carefully to obtain maximum yields and activity. In addition to the typical phenomenon of the inhibition by methanol, the cases of inhibition by both methanol and vegetable oil,[35] as well as by the reaction products such as glycerol[38] have been reported. On the other hand, the presence of water even at very low concentrations strongly favors the activity of enzymes.[39] An accurate analysis of kinetic data on the transesterification process can be very helpful for process optimization; however, the kinetic details and models describing the kinetics of specific processes can be drastically dependent on some other reaction parameters, such as the nature of the solvent, catalyst, and substrate, to be discussed further herein. Overall, the classical biodiesel production process requires a quite complex production facility and has a number of advantages and drawbacks, which will be discussed below. Intensive research and technological development are underway to improve process efficiency and economics. The scope of this section, which is dedicated to transesterification technology for biodiesel production, however, is to feature and discuss scientific rather than engineering aspects. The following critical clues to more efficient biodiesel production
ChemSusChem 2009, 2, 278 300

by transesterification are addressed in the following subchapters: 1) quality of feedstock, such as free fatty acid content and moisture content, 2) reaction conditions, such as reagents molar ratio, temperature, reaction time, and mixing intensity, and 3) choice of the type of catalyst. The latter issue will be further extended to a comparison of different catalytic systems in Section 3. 2.2. Feedstocks Any vegetable oil or animal fat can be used to prepare biodiesel, however, a careful choice of feedstock should be made. Currently, biodiesel is mostly produced from edible and nonedible vegetable oils obtained from rapeseed,[4042] soybean,[26, 4357] oil palm,[5871] coconut,[72, 73] sunflower,[7487] and Jatropha curcas.[8890] The use of oils from andiroba,[91] babassu,[72, 91] camelina,[92] cumaru,[91] cynara cardunculus,[93, 94] groundnut,[95] karanja,[96] algae,[94, 97104] poppy seed,[105] rice bran,[95, 106] rubber seed,[107, 108] sesame,[83, 109] tobacco seed,[110] palm kernel,[73, 111113] fish,[104, 114, 115] and others[116] for the preparation of biodiesel has been reported also. The source of feedstock for the production of biodiesel should fulfil two requirements: price (low feedstock and production costs; more than 80 % of the production cost corresponds to the feedstock cost) and local availability (large and constant production volume).[117] It is also necessary to take into consideration the oil content of the seeds and the yield of oil per hectare. Detailed data regarding the oil-yielding capacity of different plants has been reported, from which it emerges that oil palm, under commercial planting conditions, can give an excellent oil yield per hectare of about 4000 kg.[118] Transesterification technology for used cooking oils and animal fats is more complex as compared to fresh vegetable oils. The free fatty acids (FFAs) contained in waste oils can react with the commonly used alkaline catalysts resulting in deactivation of the catalyst and in unfavorable saponification phenomenon. The high FFA content and other low quality of vegetable oils are usually the result of their exposure to heating, moisture, and oxygen. The latter causes hydrolysis and cracking of oils to FFAs, as well as their oxidation and biodegradation. As a rule, used cooking oils, oils produced in summer,[119] and oils stored for a long time exhibit a higher FFA content. For example, the FFA content of used oil from food processing is reported to be in the range of 5 to 30 %.[15] Generally, conventional transesterification technology requires that the reagents be free of moisture and the FFA content does not exceed 0.10.5 wt %.[15, 116] The use of oils with a high amount of FFAs significantly decreases the biodiesel yield owing to FFAs reacting with the alkaline catalysts forming soaps and water, both of which must be removed during the subsequent ester purification step. Notably, the presence of soap and water also reduces the efficiency of the catalyst and renders product separation difficult. Vegetable oils with a high FFA content should be refined, otherwise they require a combined acid-catalyzed esterificationbase-catalyzed transesterification process.[119] In this case, use is made of FFAs as both reactions result in the formation of FAME biodiesel.

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

283

S. Miertus et al. It is important to point out that the properties of biodiesel are strongly influenced by the nature of the fatty acid chains in the triglyceride,[120] such as their length, degree of unsaturation, and presence of other chemical functional groups, which also strongly depends on the source of the feedstock. In general, the cetane number, cloud point, heat of combustion, melting point, and viscosity of the biodiesel increase with increasing chain length of the component fatty acid and decrease with increasing unsaturation.[121] In addition, esters with unsaturated fatty acid chains are less stable as a result of relatively fast oxidation. The melting point of vegetable oils is a key property, as the transesterification process is carried in the liquid phase and oils with a high melting point may require heating to higher temperatures, which ultimately increases the energy requirements and production costs. Furthermore, the transesterification process does not alter the fatty acid chain composition and therefore the biodiesel made from feedstocks containing higher concentrations of high-melting-point saturated longchain fatty acid chains tends to have relatively poor cold flow properties.[15] The presence of impurities and calorific value content of the feedstock also affects the quality of the biodiesel. 2.3. Influence of Reaction Time, Mixing Intensity, and Temperature It is necessary to mention some important process variables of biodiesel production from vegetable oils such as the methanol/oil ratio, reaction time, mixing intensity, and temperature. The molar ratio of alcohol to triglycerides is an important variable that affects the yield of biodiesel in the transesterification reaction. Most systems for transesterification employ an alcohol/triglyceride molar ratio of 6:1.[18] The excess of methanol with respect to the reaction stoichiometry is needed to shift the reaction equilibrium to the right (product side). Clearly, the conversion of the oil into the ester is also strongly dependent on the reaction time and temperature. Typically, the transesterification reaction is complete within around 1 h using a methanol/oil molar ratio of 6:1 at a reaction temperature of 60 8C.[15] Freedman et al.[18] compared the transesterification of peanut, cottonseed, sunflower, and soybean oil using a methanol:oil molar ratio of 6:1, 0.5 % sodium methoxide as catalyst, and a reaction temperature of 60 8C. A biodiesel yield of around 80 % was observed after already 1 min for the soybean and sunflower oils. After 1 h, a conversion of 9398 % was reached for all four oils. Even if some oils can be efficiently converted into their esters at ambient temperature, such as castor oil into methyl ricinoleate,[122] most systems are designed to operate between 60 and 70 8C.[18] Refaat et al.[123] found that the optimum temperature for the transesterification of various types of oils is 65 8C, with most reactions being complete within 30 min at this temperature, whereas at 25 8C significantly lower yields were obtained even after 1 h of reaction. In the study by Freedman et al.,[18] after 6 min of reaction the corresponding esters were obtained in yields of 94, 87, and 64 % at 60, 45, and 32 8C. In the latter case (reaction at 32 8C) nearly 4 h reaction time was required to reach 99 % yield. The rate of the transesterification reaction of vegetable oil with alkaline methanol solution strongly depends on the rate of mass transfer at the interface between glycerolmethanol and oilester phases. Generally, low reaction rates are observed in transesterification as a result of a poor dispersion of the methanol and oil phases, and an induction period can be often seen on the kinetic curves (slow initial reaction before steady-state concentrations are reached). Therefore, intense mixing is very important for the transesterification process. Noureddini et al.[124] investigated the effect of variations in mixing intensity during the transesterification of triglycerides to methyl esters in a pilot plant. The optimum stirring rates were in the range of 1000 rpm using both motionless and high-shear mixers. Singh et al.,[125] using a statistical approach, were able to optimize the operating conditions to maximize the biodiesel yield and minimize the formation of soap. The following set of optimized conditions was indicated: potassium methoxide concentration of 0.2 mol L1, a reaction temperature of 50 8C, and a methanol/oil molar ratio of 4.5:1.

3. Choice of Catalyst for Transesterification


The selection of an appropriate catalyst is of fundamental importance for the design of a sustainable transesterification proheterogenecess. Homogeneous,[69, 80, 126130] [52, 53, 73, 75, 76, 82, 87, 131135] ous, and enzymatic[18, 136, 137] catalysts have been investigated. Currently, homogeneous alkaline catalysts, such as sodium hydroxide and potassium hydroxide, are most commonly used in industrial transesterification processes for biodiesel production, mainly because they are able to efficiently promote the reaction at relatively low temperatures. Homogeneous acid catalysts and heterogeneous (solid) catalysts are used to a lesser extent. While the main advantages in the use of homogeneous acid and base catalysts are their cost-effectiveness and good performance, both catalysts require the use of an excess of alcohol, can provoke various problems, and are associated to technical difficulties. The use of homogeneous catalysts is normally limited to batch-mode processing followed by a catalyst separation step. The immiscible glycerol phase, which accumulates during the course of the reaction, solubilizes the homogeneous base catalyst and, therefore, withdraws it from the reaction medium. Moreover, other difficulties of using homogeneous base catalysts relate to their sensitivity to FFAs and water and the resulting saponification phenomenon. Usually, base catalysts are better suited to the transesterification of triglycerides, whereas acid catalysts are better suited for the esterification of FFAs. Therefore, feedstocks containing FFAs may require two different types of catalysts (acid and base catalysts) that are usually employed in a two-stage process, which means that the acid catalyst from the first stage has to be removed before the base catalyst is added in the second stage. A disadvantage of homogeneous acid catalysts is also their increased corrosiveness. Alternatively, the esterifiChemSusChem 2009, 2, 278 300

284

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production cation and transesterification can be combined in a consecutive continuous process,[99] which would require the combination of acid and base heterogeneous catalysts. Solid bifunctional catalysts that are able to effectively catalyze both processes simultaneously (one pot or in continuous mode) have been reported.[138] Another cost-effective solution for the conversion of feedstocks with high FFA content into FAMEs lies in the consecutive hydrolysisesterification process (hydroesterification). In the first step, the triglyceride content of the oil is stoichiometrically hydrolyzed into the corresponding FFAs using a well-known industrial steam hydrolysis process. The FFAs formed then undergo selective esterification over an acid catalyst.[139] Heterogeneous catalysts have the general advantage of being reusable and easy to separate from the reaction products (generally cleaner process). In addition and more specifically, they do not form soaps, are more selective to biodiesel (purer product), and simplify the glycerol purification (99 % purity glycerol can be produced against 75 % in the homogeneous process). They are in general much more tolerant to water and FFAs in the feedstock. Nevertheless, heterogeneous catalysts for biodiesel production generally require more severe working conditions than those for homogeneous catalysts, in particular, higher temperatures and pressures. Those on solid supports tend to be less active than the active species in solution. Finally, there is always the possibility of leaching, which might contaminate the biodiesel and reduce the lifetime of the catalyst. The growing attention to develop biodiesel production processes based on an enzymatic approach is justified by the use of more moderate reaction conditions, a lower alcohol-to-oil ratio, easier product recovery, and higher environmental compatibility than chemical catalysts (homogeneous or heterogeneous). Furthermore, FFAs present in oils/fats can be completely converted into alkyl esters using enzyme catalysts. Notably, new immobilization technology indicates that enzyme catalysts can become cost-effective as compared to chemical processing. However, the production cost of lipases, which is the most investigated and promising class of enzyme catalysts for transesterification, is still significantly higher than that of alkaline catalysts. Enyme-based technology is still at a stage of intensive research and process optimization. 3.1. Homogeneous Base Catalysts The transesterification of triglycerides with low-molecularweight alcohols catalyzed by homogeneous alkali catalysts is the most common process in the biodiesel industry because of the low cost of the catalyst and the good conversions that can be achieved in short reaction times and at moderate temperatures. Catalysts such as alkaline metal hydroxides,[80, 127129, 140] alkoxides,[18, 44, 125] and carbonates[82] are most often used. Commercially, NaOH and KOH are preferred because of their availability and low cost. The alkoxides are more expensive than the hydroxides and are more difficult to handle because they are hygroscopic. However, as the hydroxide ion is present only as an impurity in alkoxides,[13] they do not produce soap through
ChemSusChem 2009, 2, 278 300

triglyceride saponification. However, the general limitation for the use of homogeneous alkaline catalysts is the total FFA content associated with the feedstock, which must not exceed 0.5 wt %; otherwise, soap formation seriously hinders the production of fuel-grade biodiesel.[13] Ma et al.[128] investigated the effect of alkaline catalysts on the transesterification of beef tallow. NaOH was found to perform significantly better than NaOMe. Furthermore, a slightly higher concentration of NaOMe with respect to NaOH (0.5 vs. 0.3 % w/w) was needed to obtain the maximum conversion of the oil into the corresponding esters. Singh et al.[125] studied the reaction of methanol with canola oil at different concentrations of alkaline catalyst (NaOH, KOH, NaOMe, and KOMe), reaction temperatures, and methanol-tooil molar ratios. A preliminary analysis of the variance performed on the four process variables and their possible twoway interactions showed that only the catalyst formulation concentration and catalyst formulationreaction temperature had significant contributions among the two-way interactions. The results showed that there were significant differences in the product yields among the four catalyst formulations. Potassium-based catalysts gave better yields than the sodium-based catalysts, and methoxide catalysts gave higher yields than the corresponding hydroxide catalysts. On the other hand, potassium-based catalysts resulted in a larger extent of soap formation than the corresponding sodium-based catalysts. Notably, the cost of potassium and sodium methoxides is five to six times higher than potassium and sodium hydroxides. An increase in reaction temperature had a positive effect on both transesterification and saponification reactions. Increasing the MeOH/oil molar ratio had a positive effect on the reaction yields as it helped to drive the reaction equilibrium forward. The molar ratio of methanol to oil, however, had an unpredicted effect on the saponification. The amount of soap formed decreased with an increase in the molar ratio from 3:1 to 4.5:1 but increased when the molar ratio increased from 4.5:1 to 6:1. On the other hand, as the catalyst concentration in the methanol phase was relatively high at lower MeOH/oil ratios, higher reaction rates were observed. Potassium and sodium hydroxides and methoxides were also investigated as catalysts by Vicente et al.[130] in the transesterification of sunflower oil using a methanol-to-oil molar ratio of 6:1 and 1 % catalyst. The yields of esters were reported to be higher than 98 % for the methoxide catalysts, and 85.9 and 91.67 wt % for the sodium and potassium hydroxides, respectively, because the saponification resulted in more substantial decreases in yield. Leadbeater et al.[141] showed that the use of microwave heating allows an efficient and easy preparation of biodiesel with the advantage of a short reaction time. The process can be effectively adapted for both batch and continuous-flow preparation of biodiesel and in the presence of either homogeneous acid or base catalysts.[142, 143] Azcan et al.[144] studied the effect of microwave heating on the transesterification of cottonseed oil using NaOH and KOH as catalysts. The results indicated that microwave heating effectively reduced the reaction time from 30 min (for a conventional heating system) to 7 min.

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

285

S. Miertus et al. Tang et al.[145] studied the catalytic activity of triethylamine, diethylamine, ethylenediamine, and monoethanolamine in the transesterification reaction of rapeseed oil in supercritical methanol. The catalytic activity decreased in the order: diethylamine > triethylamine > monoethanolamine > ethylenediamine. With triethylamine as catalyst, it was observed that an increase in pressure has a positive effect on the transesterification reaction. Furthermore, the use of an appropriate co-solvent, such as ethyl acetate, can improve yields. Arzamendi et al.[82] investigated the catalytic activity and selectivity of hydroxides of Li, Na, K, Rb, Cs, and Ca in the transesterification of refined sunflower oil with methanol at 50 8C. The activity was compared to that of a heterogeneous system, such as carbonates of Na, K, Ca, and Mg, bicarbonates of N and K, and oxides of Ca and Mg. Notably, the catalytic experiments were conducted using a large excess of methanol (methanol/oil molar ratio of 12:1). The alkaline metals hydroxides were the most active catalysts, although potassium and sodium carbonates showed appreciable performances also. In the case of K2CO3, its significant solubility in methanol suggests that the contribution of homogeneous reaction routes can be important. Schuchardt et al.[146] investigated the transesterification of rapeseed oil with methanol using substituted cyclic and acyclic guanidines and compared the results with the unsubstituted guanidine. The catalytic activity of the guanidines depended mainly on their intrinsic base strength. No lipophilic effect was observed with a long alkyl chain on the guanidine. 1,5,7TriazabicycloACHTUNGRE[4.4.0]dec-5-ene showed the best performance. In fact, when used at 1 mol %, the methyl esters were obtained in 90 % yield after 1 h of reaction. Schuchardt and co-workers also investigated many other non-ionic bases, especially with the goal to make them heterogeneous.[52, 147] Very recently, Cerro et al.[148] reported the promising performances of non-ionic triaminoACHTUNGRE(imino)-phosphoranes (phosphazenes) as base catalysts for the water-free alcoholysis of vegetable fatty esters. The use of two catalysts, with different intrinsic basicity, allowed the direct correlation of the activity to the basicity of the catalyst. Very high yields (9395 % mol) of biodiesel were achieved under rather mild reaction conditions. Furthermore, the study indicated that the catalyst could be recovered and recycled. 3.2. Homogeneous Acid Catalysts Homogenous acid catalysts are significantly more effective than base catalysts in the esterification of FFAs but are not as effective as the base catalysts in the transesterification of triglycerides. Acid-catalyzed transesterification proceeds approximately 4000 times slower than that catalyzed by the same amount of an alkali catalyst[17] and also needs harsher temperature and pressure conditions.[149] However, the advantages of acid catalysts include their low sensitivity to moisture and absence of soap formation. Acid catalysts can be used in the transesterification of feedstocks containing a high content of FFAs, such as waste cooking oils, where the use of base catalysts is undesirable because of the formation of soap. Acid catalysts can be used also in a two-stage process, in which the first stage involves the esterification of FFAs into biodiesel in the presence of the acid catalyst followed by base-catalyzed transesterification. The transesterification of different triglycerides in the presence of hydrochloric, sulfuric, phosphoric, and organic sulfonic acids have been extensively studied.[44, 74, 149] Freedman et al.[44] investigated the conversion of soybean oil into its methyl ester in the presence of 1 % sulfuric acid with an alcohol/oil ratio of 30:1 at 65 8C and observed that a reaction time of 50 h was required to obtain yields over 99 %. Butanolysis at 117 8C and ethanolysis at 78 8C using the same quantities of catalyst and alcohol required 3 and 18 h, respectively. Different types of feedstocks, such as rice bran oil[106] and waste cooking oil,[150] have been successfully transesterified with methanol using sulfuric acid as catalyst in the temperature range 6080 8C. In particular, Zullaikah et al.[106] investigated the acid-catalyzed methanolysis of dewaxed/degummed rice bran oil with varying FFA contents at atmospheric pressure and 60 8C using 1:10 molar ratio of oil/methanol and 2 wt % sulfuric acid as catalyst. The initial FFA content appreciably influences the rate of methanolysis and the final methyl ester content in the product. A methyl ester content of about 96 % in the product could be obtained in 8 h for rice bran oil with an initial FFA content of 76 %. Notably, analysis of the reaction products showed constant residual acid content (23 wt %), which might be attributed to the presence of phenolic compounds such as oryzanol, tocols, and/or other minor components in the rice bran oil. Consistent with the fact that acid-catalyzed alcoholysis of triglycerides is a slow reaction, a substantial amount of unreacted triglyceride was detected even after 24 h if the starting oil contained more than 20 % FFAs. A further increase in reaction time or methanol concentration had a negligible effect on the transesterification of the triglycerides. To overcome these problems, a two-step acid-catalyzed methanolysis process was adopted to obtain good conversions in reasonably short times using rice bran oil with a high FFA content (> 20 %) as feedstock.[106] The first step was performed at 60 8C using 2 wt % sulfuric acid and an oil/methanol molar ratio of 1:5. At a reaction time of 2 h, when more than 98 % of the FFAs were converted into the corresponding methyl esters, the organic phase containing the reaction product was removed and used as the substrate for the second methanolysis reaction at 100 8C. At the end of the first step, the reaction product contained 62 % FAMEs, 3.2 % FFAs, and 34.8 % acylglycerides. Methanolysis of this product at 100 8C resulted in a FAME content of more than 96 % in the final mixture for a total reaction time of 8 h. Zheng et al.[150] showed that, in a large excess of methanol, the acid-catalyzed transesterification reaction of waste cooking oils is essentially a pseudo-first-order reaction. The oil/methanol/acid molar ratio and temperature were the most significant factors affecting the yield of FAMEs. At 70 8C with an oil/methanol/acid molar ratio of 1:245:3.8, and at 80 8C with oil/methanol/acid molar ratios in the range from 1:74:1.9 to 1:245:3.8, the transesterification was essentially a pseudo-first-order reaction as a result of the large excess of methanol which drove
ChemSusChem 2009, 2, 278 300

286

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production the reaction to completion (99 % at 4 h). In the presence of the large excess of methanol, the FFAs present in the waste cooking oil were very rapidly converted into methyl esters in the first few minutes under the above conditions. Little or no monoglycerides were detected during the course of the reaction, and the diglycerides present in the initial waste cooking oil were rapidly converted into FAMEs. Recently, Aranda et al.[151] investigated the acid-catalyzed homogeneous esterification of palm fatty acids with anhydrous and hydrated methanol and ethanol. With anhydrous methanol, it was noted that sulfuric and methanesulfonic acids were the best catalysts, with conversions higher than 90 % obtained after 1 h reaction at 130 8C. The esterification reaction using anhydrous ethanol showed similar results. On the other hand, trichloroacetic acid did not accelerate the reaction, and phosphoric acid yielded minor improvements in conversion. Notably, even a small amount of sulfuric acid (0.01 % w/w) proved to appreciably promote the reaction. A water inhibition effect was observed mainly in the ethanolic reaction. Abreu et al.[43, 91] studied the transesterification of different types of Brazilian native oils using organometallic catalysts such as Sn(3-hydroxy-2-methyl-4-pyrone)2ACHTUNGRE(H2O)2, Pb(3-hydroxy2-methyl-4-pyrone)2ACHTUNGRE(H2O)2, and Zn(3-hydroxy-2-methyl-4pyrone)2ACHTUNGRE(H2O)2 at alcohol/vegetable oil/catalyst molar ratios of 400:100:1. These bivalent metal complexes can homogeneously catalyze the transesterification reaction by acting as Lewis acid sites. It was noted that the tin complex was the most active for all the vegetable oils tested. Furthermore, the catalysts showed an increased activity when vegetable oils with short-chain fatty acids or with a high degree of unsaturation and short linear alcohols were used. cerides and to avoid deactivation of catalytic sites by strong adsorption of polar by-products such as glycerol and water. As in the case of homogeneous catalysts, microwaves have shown superior performance than traditional heating methods for biodiesel production, with heterogeneous catalysis leading to better yields and conversions of triglycerides into FAMEs in a short time and, consequently, less energy consumption.[152, 153] One of the main problems with heterogeneous catalysts is their deactivation with time owing to many possible phenomena, such as poisoning, coking, sintering, and leaching. The problem of poisoning is particularly evident when the process involves used oils. More general and dramatic is catalyst leaching, which not only can increase the operational cost as a result of replacing the catalyst but also leads to product contamination. While purification of the reaction substrates, optimization of process parameters, and inclusion of catalyst-regeneration steps can minimize the deactivation of heterogeneous catalysts, leaching phenomena can be minmized only by altering the catalyst formulation. In particular, it is essential to use robust materials that are able to resist attrition. Furthermore, it is important to enhance the interaction between the active phase and the support. This enhancement can be obtained by tuning the synthesis parameters, for instance, by designing embedded catalysts. Finally, partial dissolution of the catalyst opens up the serious possibility of co-existing homogeneously catalyzed reaction pathways. Several heterogeneous catalysts have been reported for the transesterification of vegetable oils (see Table 2).

3.3.1. Acid Zeolites Among numerous other applications, zeolites are probably the most investigated inorganic solid acid catalysts for the production of biodiesel by transesterification, as these materials can be synthesized with extensive control of acidic/basic and textural properties. The acidic properties of zeolites are usually improved by protonation, that is, by exchange of the cations contained in the positively charged aluminosilicate cage with protons. Furthermore, it is possible to induce some hydrophobicity of zeolites by the elimination of water of hydration. Koh et al.[154] studied the transesterification of waste cooking oil using various zeolite catalysts with different acidities and pore structures. Proton-exchanged MOR, MFI, FAU, and BEA zeolites were employed in the reaction with silicalite, which has no strong acid sites. The yields were found to be independent of the pore structure of the zeolites. The authors found that the yield increased linearly with enhancing the acid strength and with increasing the amount of strong acid sites. Consistently, proton-exchanged MOR(10) zeolite, which has more acid sites and greater acid strength than other zeolites, induced the highest yield of methyl esters (95 %). The stability of zeolites, especially at moderate/high temperatures and in the presence of water, must be carefully considered. In fact, the consequent dealumination process can lead to a serious decrease in the quantity of acid sites and the acid strength, with consequent deactivation of the catalyst.

3.3. Heterogeneous Catalysts High energy consumption and costly separation of the catalyst from the reaction mixture have inspired the development of heterogeneous catalysts. The use of heterogeneous catalysts does not lead to the formation of soaps through neutralization of FFAs or saponification of triglycerides. Furthermore, solid acid catalysts can indeed improve the sustainability of the biodiesel production process, eliminating the corrosion problems associated with their use and consequent environmental hazards posed by them. However, the performance of heterogeneous catalysts is generally lower than that of the commonly used homogeneous catalysts. Moreover, heterogeneously catalyzed transesterifications require relatively elevated temperatures and pressures. Notably, diffusional limitations might sometimes drastically reduce the surface of the solid that is available for promoting the transesterification reaction. Therefore, a careful design of the pore structure of these materials is important. In this respect, zeolites are ideal systems. Furthermore, to improve the performance of these catalysts, it is essential to understand the correlations between acid and base strength and catalytic activity. It is clear that the surface of these heterogeneous materials should display some hydrophobic character to promote the preferential adsorption of triglyChemSusChem 2009, 2, 278 300

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

287

S. Miertus et al.

Table 2. Main classes of solid base and acid catalysts employed for transesterification reactions. Heterogeneous acid catalysts Protonated Y zeolites (HY): H-BEA, H-ZSM-5, H-MOR, H-MFI, H-FAU, etc. Keggin heteropolyacids: HnXM12O40 (X = P, Si; M = Mo, W) and their salts (can be supported on ZrO2, TiO2, etc.) Mixed metal oxides (e.g. ZrO2 and SnO sulfated and/or doped with Al, Ti, W, Si, or alkali metals) Sulfonic acid linked to a polymer framework (ion-exchange resins, e.g. Amberlyst-15, Nafion) or immobilized (e.g. organosulfonic acids on silica) Fe-Zn double-metal cyanide complexes Heterogeneous base catalysts Zeolites exchanged with strongly basic cations (e.g. Cs, K) or containing other occluded basic species Hydrotalcites of Mg-Al or Li-Al: [Mg(1x)Alx(OH)2]x + ACHTUNGRE(CO3)x/n2 Metal oxides: CaO, MgO, La2O3, ZnO, etc. Insoluble basic salts/hydroxides: species loaded on alumina (e.g. KI, KF, K2CO3, Na/NaOH), carbonates, etc. Organic bases: guanidines immobilized on polymers, salts of amino acids containing guanidino or quaternary ammonium groups, etc. Basic oxides supported on high-surface-area materials (e.g. SBA-15, MCM-41) Alkali earth alkoxides (e.g. CaACHTUNGRE(OMe)2)

Shu et al.[155] prepared a La/zeolite beta catalyst by the ionexchange method and used it in the batch transesterification of soybean oil with methanol. The prepared La/zeolite beta catalyst showed higher conversions and better stability than zeolite beta which was correlated to the larger quantity of external Brnsted acid sites. 3.3.2. Heteropolyacids Heteropolyacids (HPAs) have been the subject of recent attention as a result of their excellent water tolerance, superacidity, and porous architecture. Notably, HPAs can be slightly soluble in the reaction media and, therefore, the homogeneous reaction can contribute to the overall reaction rate. This can be also true for other supported systems discussed here in which leaching phenomena are significant. Chai et al.[156] investigated the use of Cs2.5H0.5PW12O40 as a heterogeneous catalyst for the production of biodiesel from Eruca sativa Gars. oils (ESG oil) with methanol. They reported that this HPA shows almost the same activity under optimized reaction conditions as a conventional homogeneous catalyst such as sodium hydroxide or sulfuric acid, but with the advantages of easy separation and catalyst reuse. The advantages claimed for the use of this catalyst system are that the reaction is feasible at room temperature, the catalytic activity is not affected by the content of FFAs or water in the feed, and the reaction times are shorter. Even though the use of supported HPAs was also suggested, note that these systems may be more hydrophilic and hence may not be water-tolerant. It is necessary that the activity of this class of catalysts be evaluated with respect to other hydrophobic solid catalysts especially from the point of variables such as the methanol-to-oil molar ratio, reaction temperature, and the extent of reusability. Xu et al.[157] investigated the use of a mesoporous polyoxometalatetantalum pentoxide composite catalyst (H3PW12O40/ Ta2O5). Different H3PW12O40 loadings, ranging from 3.6 to 20.1 %, were obtained by a one-step sol-gel hydrothermal route in the presence of a triblock copolymer surfactant. The composite exhibited large and well-distributed three-dimensionally interconnected pores (3.95.0 nm), a high BET surface area (around 110 m2 g1), good porosity (0.441.37 cm3 g 1),

and a homogeneous dispersion of the Keggin unit throughout the composite. The system showed a promising catalytic performance in the esterification of lauric acid and myristic acid, in the transesterification of tripalmitin, as well as in the transesterification of soybean oil. In the presence of FFAs, H3PW12O40/ Ta2O5 showed high activity and selectivity towards simultaneous esterification and transesterification under mild conditions. Finally, the catalyst was recovered, reactivated, and reused several times, thus opening perspectives for its application. Cao et al.[158] studied the use of the heteropolyacid H3PW12O406 H2O in the transesterification of waste cooking oils with high acid and water contents. The hexahydrate polyacid was found to be a very promising catalyst: in the transesterification of waste cooking oil with methanol the ester was obtained in 87 % yield, whereas in the esterification of long-chain palmitic acid the ester was obtained in 97 % yield. However, a reaction time of at least 10 h was required. The catalyst could be separated from the products by distillation of the excess methanol and was reused several more times. The catalytic activity was not affected by the FFA content and water content in the waste cooking oil, and the transesterification occured at a lower temperature (65 8C) and at a lower methanol/oil ratio (70:1). Very recently, Alsalme et al.[159] investigated the esterification of hexanoic acid and the transesterification of ethyl propanoate and ethyl hexanoate with excess methanol (1:20 molar ratio) at 60 8C and ambient pressure with a range of HPA catalysts. The turnover frequencies decreased with decreasing acid strength of the catalyst in the order: H3PW12O40 % Cs2.5H0.5PW12O40 > H4SiW12O40 > 15 % H3PW12O40/Nb2O5, 15 % H3PW12O40/ZrO2, 15 % H3PW12O40/TiO2. Bulk cesium salt Cs2.5H0.5PW12O40 all exhibited a high catalytic activity as well as a high stability to leaching. The authors clearly highlighted that the supported HPA catalysts suffer from leaching and that the reaction exhibits a significant contribution from homogeneously catalyzed pathways. 3.3.3. Sulfated Zirconia and Mixed Metal Oxides Sulfated zirconia catalysts have been extensively examined in the transesterification reactions of vegetable oils as a result of
ChemSusChem 2009, 2, 278 300

288

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production their good acidity.[160, 161] Notably, advanced preparation protocols involving the use of chlorosulfonic acid as a source of sulfate ion can induce superacidity.[162] The major disadvantage of these systems is the leaching of sulfate ions during operation. Therefore, the contribution of homogenous catalysis cannot be neglected.[163] As an alternative to sulfated zirconia, sulfated tin oxide as well as tungstated zirconia-alumina have been examined also.[164, 165] It has been suggested that the presence of monoclinic zirconia and crystalline WO3 could have detrimental effects and that active sites are generated only at the interface between tetragonal zirconia and amorphous WO3. Kiss et al.[160] investigated various zeolites, ion-exchange resins, and mixed metal oxides for the esterification of dodecanoic acid with 2-ethylhexanol, 1-propanol, and methanol at 130180 8C. The study confirmed that the most promising catalyst was sulfated zirconia. Jitputti et al.[166] showed that sulfated zirconia exhibited a good activity in transesterification processes of palm kernel oil and coconut oil. In fact, the use of 1 wt % catalyst at 200 8C resulted in a FAME content higher than 90 % for both substrates. Furuta et al.[165] studied tungstated zirconia-alumina, sulfated tin oxide, and sulfated zirconia-alumina as catalysts for the transesterification of soybean oil. Both the transesterification of soybean oil with methanol and esterification of n-octanoic acid with methanol to methyl n-octanoate were carried out in flow reactors. In the transesterification process caried out at over 250 8C, tungstated zirconia-alumina gave yields over 90 %, whereas the other two catalysts gave yields of less than 80 %. In the esterification reaction carried out at over 175 8C, all three catalysts showed quite high activities (yields close to 100 %). Sulfated tin oxide showed an especially high activity in the esterification as a result of its strong acidity. Tungstated zirconia-alumina was indicated as suitable for both reactions. Furuta and co-workers extended their investigations to titanium-doped zirconia (TiO2/ZrO2 with 11 wt % Ti), aluminumdoped zirconia (Al2O3/ZrO2 with 2.6 wt % Al), and potassiumdoped zirconia (K2O/ZrO2 with 3.3 wt % K) in the transesterification of soybean oil with methanol and in the esterification of n-octanoic acid with methanol. In continuous-flow mode, K2O/ ZrO2 gave a conversion close 100 % at the beginning of the reaction but deactivated rapidly owing to the leaching of potassium. TiO2/ZrO2 and Al2O3/ZrO2 led to yields of 100 % at 175 8C after 20 h. The authors claimed that the good performance of these catalysts in the transesterification reaction was based on their high activity in the esterification of the liberated FFAs. Du et al.[167] prepared several sulfated silica-doped tin oxides with large surface areas by hydrothermal synthesis, followed by sulfation and calcination. Catalytic tests in the transesterification of triacetin as a model substrate with methanol showed that sulfated silica-doped tin oxides were much more active than conventional sulfated tin oxides. Using 10 mmol triacetin, 90 mmol methanol, and 100 mg of sulfated silica-doped tin oxide (Si/Sn molar ratio = 0.20) at 60 8C, a conversion of 91.1 % and a selectivity to glycerol of 68.4 % were obtained after 12 h. For comparison, under the same conditions, standard sulfated zirconia and standard sulfated tin oxide gave conversions of 66.0 and 77.5 % and selectivities to glycerol of 29.8 and 50.4 %,
ChemSusChem 2009, 2, 278 300

respectively. Notably, the presence of SiO2 seemed to have a positive effect on the lifetime of the catalyst. 3.3.4. Ion-Exchange Resins and Immobilized Sulfonic Acids Heterogeneous solid catalysts for the effective production of biodiesel from waste oils can be prepared by the incomplete carbonization of carbohydrates, such as d-glucose, sucrose, cellulose, and starch, followed by sulfonation.[168, 169] These catalysts are highly effective, can be recycled, and are compatible with high-acid-value substrates. In fact, the carbohydrate-derived catalysts exhibited substantially higher catalytic activities for both esterification and transesterification than sulfated zirconia and niobic acid and gave markedly enhanced yields of methyl esters in converting waste cooking oils containing 27.8 wt % FFAs to biodiesel. In addition, under optimized reaction conditions, the starch-derived catalyst retained a remarkably high proportion (about 93 %) of its original catalytic activity even after 50 cycles of reuse. Shibasaki-Kitakawa et al.[170] tested the catalytic performances of various ion-exchange resin catalysts, such as PK208, PA308, PA306, PA306s, and HPA25, in the transesterification reactions of triolein with ethanol to produce ethyl oleate. Anionexchange resins exhibited appreciably higher catalytic activities than the cation-exchange resins. Furthermore, the anion-exchange resins with a lower cross-linking density and a smaller particle size led to a high reaction rate as well as to a better conversion. Finally, the possibility of a simple means to regenerate the catalyst is promising wih respect to a continuous transesterification process. Lopez et al.[133] investigated different acid and base solid catalysts, also in comparison with NaOH and H2SO4, in the transesterification of triacetin (as a model substrate) with methanol at 60 8C in a batch reactor. Among various known systems, Amberlyst-15 and Nafion NR50 showed reasonable activities. More specifically, Dos Reis et al.[72] investigated the preparation of biodiesel from vegetable oil using commercially available sulfonic ion-exchange polymeric resins. Amberlyst-15 showed the best performances among the investigated commercial resins. Notably, these systems showed better activity than homogeneous catalysts. The transesterification of soybean oil with methanol at 60 8C and atmospheric pressure was studied by Guerreiro et al.[46] in the presence of nafion membranes, ion-exchange resins, and poly(vinyl alcohol) membranes containing sulfonic groups. Nafion and PVA membranes in the form of films were also used in a membrane reactor. The PVA membrane modified with sulfosuccinic acid was, in both cases, the most active catalyst. Mbaraka et al.[171173] studied the performance of organosulfonic acid incorporated in mesoporous silica materials, also further functionalized with hydrophobic organic groups, in the esterification of fatty acids with methanol. It was found that the acid strength was a function of the spatial location of the organosulfonic acid groups on the surface. A shorter distance led to an increase in the acid strength and a better performance of the catalyst, as a result of the interaction between

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

289

S. Miertus et al. neighboring acid sites. The most acidic of the prepared silicas, SBA-15-ArSO3H, however, resulted in only 47 % conversion of palmitic acid in the esterification with methanol (1:20 acid/ methanol ratio) after 30 min at 85 8C.[172] 3.3.5. Double-Metal Cyanide Complexes Fe-Zn double-metal cyanide complexes are relatively cheap and effective catalysts for the ring-opening polymerization of propylene oxide[174] as well as the coupling of epoxides and CO2 for the production of polycarbonates.[175] The Lewis acidic Zn ions represent highly active sites for the transesterification reaction.[176] Recently, Sreeprasanth et al.[177] reported the application of Fe-Zn double-metal cyanide complexes as solid catalysts in the preparation of fatty acid alkyl esters from vegetable oils. The hydrophobic nature of these catalysts and the presence of Lewis acidic sites lead to their high activity in the simultaneous transesterification of triglycerides and esterification of FFAs. Notably, Brnsted acid sites and basic sites are absent in these systems. As a relationship between the transesterification activity and the concentration of strong Lewis acid sites has been observed, most of the attention has been dedicated to enhance the concentration of coordinatively unsaturated Zn2 + ions in the structure by modifying the preparation procedure. In particular, while these catalysts can be easily prepared from aqueous solutions of ZnCl2 and K4Fe(CN)6, it is essential to use an adequate complexing agent such as tert-butyl alcohol and a co-complexing agent such as the triblock copolymer EO20PO70EO20 (average molecular weight: 5800 g mol1).[176] 3.3.6. Base Zeolites Zeolites are flexible and versatile catalytic materials, whose acidity and/or basicity can be modulated by appropriate doping. While some zeolites can exhibit acidic properties (see Section 3.3.1), the alkali-ion-exchanged zeolites show good basicity. The basicity increases with an increase in the electropositive nature of the exchanged cation. Furthermore, the occlusion of alkali metal oxide clusters in zeolite cages, which can be obtained by the decomposition of impregnated alkali metal salts, leads to a further increase in the basicity of these materials. Corma et al.[178] investigated the glycerolysis of triolein and rapeseed oil using base catalysts, such as Cs-MCM-41, Cs-Sepiolite, MgO, and calcined hydrotalcites with different Al/Al + Mg ratios. A wide range of basicities was covered in order to extract active hydrogen atoms with pKa values from 9 to 16. MgO and hydrotalcites with a low Al content were basic enough to extract protons from aliphatic alcohols such as glycerol and therefore were active and selective catalysts for glycerolysis. Suppes et al.[53] investigated the transesterification of soybean oil with methanol in the presence of NaX faujasite zeolites, ETS-10 zeolites, zeolites exchanged with potassium and cesium, NaX faujasite zeolite containing occluded sodium oxide and sodium azide, as well as a series of metal catalysts (oxides and carbonates of Ba, Ca, Zn), and several metals (Ni, Pd, Fe). The study indicated that ETS-10 catalysts provided significantly higher conversions than the zeolite-X type catalysts. Increased conversions were attributed to the higher basicity of the ETS-10 zeolites and their larger pore structures, which improved intraparticle diffusion. Notably, the ETS-10 catalyst provided 32 % conversion of the oil to methyl esters at room temperature in 24 h. The yield of methyl esters increased with an increase in temperature. A preliminary stability test was conducted in the case of the NaX containing occluded sodium oxide (NaOx/NaX) catalyst. After the first transesterification experiment, the reaction mixture was centrifuged and separated to provide wet solid catalysts and liquid product fractions. The catalyst was subsequently reused under the same experimental conditions to the initial alcoholysis reaction, whereby some leaching of the sodium contained in the catalyst was observed. Ramos et al.[75] carried out the transesterification of sunflower oil over mordenite, zeolites beta and X incorporated with metals (using both impregnation and ion-exchange methods) with different metal loadings. Zeolite X was agglomerated with a binder, sodium bentonite, to study how the presence of a binder could change the catalytic performance. The activity of the various catalysts was correlated with the sodium loading, which led to an increase of basicity. In fact, the conversion of sunflower oil (methyl ester content) at 60 8C increased considerably, showing a linear trend and reached 95.1 wt % for NaX occluded with three excess Na atoms per supercage. Deactivation studies were carried out for 3NaX, the most active heterogeneous catalyst in the series. This catalyst exhibited a significant drop in activity from 95.1 % yield of methyl esters in the first cycle (after 7 h reaction time) to 30.3 % yield in the second cycle (after 7 h), whilst only 4.7 % conversion was observed in the third cycle (after 7 h). After each cycle, the catalyst was washed with methanol and dried overnight at 70 8C. These facts support the hypothesis of a homogeneous-like mechanism suggesting that alkali methoxide species leach out from the catalyst. Leclercq et al.[41] studied the transesterification of rapeseed oil with methanol over cesium-exchanged NaX faujasites. Notably, a long reaction time and a high methanol-to-oil ratio was required to both achieve a high conversion of the oil and to obtain high yields of the methyl esters. 3.3.7. Hydrotalcites Hydrotalcites and hydrotalcite-like compounds are layered double hydroxides of the general molecular formula [(Mx2 + )ACHTUNGRE(My3+)(OH)2 (x + y)]Ay/nnm H2O, where M2 + and M3 + respresent divalent and trivalent metal ions, respectively, and An is an intercalated anion. Structurally, these compounds contain brucitelike[Mg(OH)2] sheets where the isomorphous substitution of Mg2 + by a trivalent cation, such as Al3 + , occurs. The resulting excess of positive charge in the layered network is compensated by the anions, which occupy the interlayer space along with water molecules.[179] Layered double hydroxides can be converted into mixed oxides with high specific surface areas and strong Lewis basic sites through controlled thermal decomposition. In these systems, strong and medium basicity
ChemSusChem 2009, 2, 278 300

290

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production can be achieved by the presence of O2 and O species, respectively.[180] The presence of Al3 + alters the acidbase properties. Therefore, the acidbase properties and related catalytic activity of these materials can be tuned by varying the Al ratio in the framework. Hydrotalcites of Mg-Al were efficient catalysts for the conversion of soybean oil into the corresponding methyl ester with a loading of 7.5 % catalyst and a methanol-to-oil molar ratio of 15:1.[181] It was observed that the ester yield was 67 % after 9 h reaction. Cantrell et al.[182] investigated a series of [Mg(1x)Alx(OH)2]x + ACHTUNGRE(CO3)x/n2 hydrotalcites with x ranging from 0.25 to 0.55, which were synthesized using an alkali-free co-precipitation route. An increase in Mg content progressively reduced the surface area of the resulting hydrotalcites. All materials were effective catalysts in the liquid-phase transesterification of glyceryl tributyrate with methanol for biodiesel production. Conversion of the triglyceride to the methyl ester occurred immediately without an induction period with the concomitant formation of the diglyceride. The yield of these primary products increased linearly with triglyceride consumption over the first 15 min of reaction, after which time the secondary transesterification of the diglyceride to monoglyceride was initiated, showing the consecutive nature of the esterification reaction. The rate increased steadily with Mg content: the Mg-rich Mg2.93Al catalyst was an order of magnitude more active than MgO, and pure Al2O3 was completely inert. The rate of reaction was correlated with interlayer electron density, which can be associated with increased basicity. Corma et al.[183] investigated the possible use of a Lewis basic hydrotalcite for the transesterification of oleic acid methyl ester with glycerol. They showed that rehydrated AlMg mixed oxides, which exhibit Brnsted basic sites, were efficient catalysts for the transesterification of FAMEs with glycerol, with excellent selectivity to monoglycerides. This selectivity was higher than that obtained with a large variety of solid Lewis base catalysts. 3.3.8. Basic Metal Oxides Alkaline earth oxides can act as solid base catalysts in the transesterification of triglycerides. Among these, Ca bases are the most promising as they are inexpensive, exhibit low solubility in methanol, and are the least toxic. Leclercq et al.[41] showed that the transesterification of rapeseed oil with methanol could be efficiently performed over a high-surface-area (300 m2 g) magnesium oxide to give a high conversion of the vegetable oil and a high yield of methyl esters, particularly when the hydroxide precursor was calcined at 550 8C. In addition, preliminary results with barium hydroxide showed a high activity of the catalyst and a high yield of esters was obtained. This catalyst was particularly effective: at a methanol-to-oil ratio of 6 and under reflux of methanol, a conversion of about 80 % of the oil was observed after 1 h, and the ester was obtained in near-quantitative yield. Reddy et al.[184] showed that nanocrystalline CaO was an efficient catalyst in the transesterification of soybean oil owing to
ChemSusChem 2009, 2, 278 300

the high surface area associated with the presence of small crystallites and lattice defects. Kawashima et al.[185] studied the catalytic activity of calcium oxide for the transesterification of rapeseed oil with methanol. The authors highlighted the need for an optimal catalyst activation protocol. In fact, by pretreating CaO with methanol, a small amount of CaO was converted into CaACHTUNGRE(OCH3)2, which acted as the initiating reagent for transesterification. Subsequently, the calcium-glycerol complex formed in the reaction of CaO with glycerol functioned as the main catalyst in transesterification. Under optimal reaction conditions (0.1 g CaO, 3.9 g methanol, 15 g oil, and 1.5 h activation time at room temperature), the methyl ester was obtained in approximately 90 % yield within 3 h at 60 8C. Xie et al.[186] used ZnO loaded with KF as a solid base catalyst in the transesterification of soybean oil with methanol which showed good activity. In this case, the activity of the catalyst correlated well with their basicity. The best performances were observed with a KF loading of 15 wt %. Xie and co-workers also investigated the transesterification of soybean oil using a Ba-ZnO catalyst.[55] The results showed that with an increase in the calcination temperature from 400 to 600 8C, the basicity of the catalysts increased which resulted in an improvement in the conversion. However, at calcination temperatures higher than 600 8C, there was a decrease both in the basicity of the Ba-ZnO catalysts as well as in the conversion of the soybean oil. 3.3.9. Insoluble/Immobilized Metal Salts and Hydroxides The catalytic activity of basic alkali-metal-containing species immobilized on alumina has also been the subject of studies focused on the transesterification of vegetable oils. For example, alumina loaded with KI was shown to be an active solid base catalyst.[54] The optimal KI loading was found to be 35 wt %, which ensured a good basicity and, therefore, a good catalytic activity. Under optimized reaction conditions, a conversion of 96 % was achieved. Kim et al.[132] investigated the transesterification of soybean oil using Na/NaOH-Al2O3 catalysts, which showed rather similar activity to a conventional homogeneous NaOH catalyst. Using the common methanol/oil molar ratio of 6:1, the yields of biodiesel obtained after 1 h were slightly below 80 % and slightly above 90 % for the heterogeneous and homogenous catalytic systems, respectively. However, at a methanol-to-oil molar ratio of 9:1 and using n-hexane as a co-solvent, biodiesel was obtained in over 90 % yield after 2 h at 60 8C with Na/NaOH-Al2O3. Ebiura et al.[187] reported that selective transesterification of triolein with methanol could be achieved at around 60 8C using alumina loaded with alkali-metal salts as a solid base catalyst. The catalytic activities were shown to be relatively insensitive to the presence of water. The most favorable performance was observed with K2CO3 on alumina, with which methyl oleate and glycerol were obtained in up to 94 and 89 % yield, respectively, after 1 h reaction at 60 8C. Suppes et al.[188] showed that a satisfactory performance of calcium carbonate catalysts in selected alcoholysis reactions

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

291

S. Miertus et al. was achieved even at high temperatures (over 200 8C), where complete conversion was observed after reaction times as short as 18 min. No decrease in the activity of calcium carbonate was noted after several weeks of utilization. 3.3.11. Alkali Earth Alkoxides Gryglewicz showed that alkaline earth metal alkoxides can catalyze the transesterification reaction.[40] Liu et al.[191] investigated the transesterification of rapeseed oil to biodiesel with methanol using calcium methoxide as a solid base catalyst. Calcium methoxide exhibits strong basicity and good thermal stability. Furthermore, calcium methoxide displayed excellent catalytic activity and stability in the transesterification of soybean oil to biodiesel with methanol, under optimal conditions of 1:1 (v/v) ratio of methanol to oil, addition of 2 % CaACHTUNGRE(OCH3)2 catalyst, 65 8C, and a reaction time of about 2 h.[49] Calcium methoxide has a moderate surface area, relatively broad particle size distribution, narrower pore size distribution, strong basicity, long catalyst lifetime, and good stability in organic solvents. A recycling experiment revealed a long lifetime of the catalyst, which maintained its activity even after being reused for 20 cycles. Calcium ethoxide has also been proposed as a catalyst for the transesterification of soybean oil with methanol and ethanol.[48] The results showed that the optimum conditions were 12:1 molar ratio of methanol to oil, 3 % CaACHTUNGRE(OCH2CH3)2 catalyst, and a reaction temperature of 65 8C. A biodiesel yield of 95.0 % was obtained within 1.5 h under these conditions. Recently, Martyanov et al.[192] investigated the transesterification of tributyrine in the presence of calcium methoxide as a solid catalyst. The observed kinetic curves pointed at a sequential reaction process. The reaction started significantly slower than with homogeneous sodium and magnesium methoxides, but then the rate increased to surpass that observed with magnesium methoxide. Even though the reaction in the presence of CaACHTUNGRE(OCH3)2 appeared to be a typical heterogeneous process, a careful investigation indicated a considerable contribution of the homogenous reaction. The loss of calcium methoxide during the reaction followed by its washing and drying was close to 35 wt %. In addition, the pH of the liquid surrounding CaACHTUNGRE(OCH3)2 at the end of the first run was pH 1213. This liquid part of the reaction mixture recovered after the first run was by itself catalytically active for the transesterification of tributyrin. The authors correlated the interaction between CaACHTUNGRE(OCH3)2 and glycerol as a major factor governing the formation of catalytically active homogeneous species. The authors also pointed at the weak heterogeneous catalytic activity of MgOxACHTUNGRE(OCH3)22 x, where a fast deactivation of the catalyst was observed owing to accumulation of the butyric salt species on the catalyst surface.[192] 3.4. Enzymes Some problems associated with conventional homogeneous catalytic processes, such as removal of glycerol and the catalyst, high energy requirements, and the need to pretreat feedstocks containing FFAs or to post-treat large amounts of waste water, can be overcome by using enzymes. Enzymatic catalysts such as lipases are able to effectively catalyze the transesterification of triglycerides with high selectivity to FAMEs either in aqueous or in non-aqueous systems.[137] Several examples of
ChemSusChem 2009, 2, 278 300

3.3.9. Solid Organic Bases Immobilized organic bases, such as guanidines anchored on polymers or encapsulated in a zeolite cage, can also act as catalysts for the transesterification of vegetable oils.[147] A number of supported guanidines have shown turnover frequencies of 1215 h1 at 59 cycles of repeated use in the transesterification of soybean oil with MeOH (oil/methanol ratio: 1:7) at 70 8C.[52] Encapsulated bases were not very efficient owing to diffusion limitations to the transfer of bulky triglycerides. Salts of amino acids that are insoluble in monovalent alcohols, glycerol, and fatty acids esters were investigated as catalysts for the methanolysis of triglycerides.[189] It was shown that Ni, Cu, Cd, La, Fe, and Zn salts of 2-amino-5-guanidinovaleric acid, carnitine, betaine, and taurine, and especially those containing quaternary ammonium or guanidino groups, displayed good catalytic activity. For example, in the methanolysis of palm oil with zinc arginate as a catalyst, nearly 90 % yields of the ester were achieved within 200 min at 135 8C and 5 bar.[162]

3.3.10. Basic Oxides Supported on High-Surface-Area Materials Basic catalysts supported on mesoporous solids have been prepared and effectively used as heterogeneous catalysts for the transesterification of ethyl butyrate, sunflower oil, and castor oil with methanol.[84] These catalysts are based on calcium oxide supported on porous silica (SBA-15, MCM-41, and fumed silica). As the active base component is distributed on the surface of the support, the optimal metal oxide loading must be tuned according to the characteristics of the support, in particular, its surface area and porosity. Thus, a sample containing 14 wt % CaO supported on SBA-15 was the most active catalyst. Remarkably, a good interaction between the active phase and support minimized problems of lixiviation. In contrast, unsupported CaO suffers from this disadvantage. Li et al.[190] investigated different mesoporous silicas loaded with MgO as solid base catalysts for the transesterification of a blended vegetable oil with ethanol to produce biodiesel. Although the study investigated only two cases, it suggested minor or no effects of the precursor salt (magnesium acetate and nitrate) on the performance of the catalyst. In contrast, the catalytic activity was affected by the type of silica support (MCM-41, KIT-6, and SBA-15) and by the loading method (impregnation and in situ coating). MgO-impregnated SBA-15 exhibited the highest activity for the production of biodiesel, with a conversion as high as 96 % obtained within 5 h. Preliminary investigations suggested that the catalyst activity was related to the surface Mg concentration.

292

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production the lipase-catalyzed production of biodiesel were reported using different feedstocks,[14] namely soybean oil, sunflower oil, palm oil, kernel and coconut oil, rice bran oil, a mixture of vegetable oils, grease, and tallow. It has been shown that the enzymatic production of biodiesel is possible by using either extracellular or intracellular lipases.[137] The choice of the method should be based on the balance between simplified upstream operations (intracellular) and high conversions (extracellular). Both types can be immobilized for use without a need for downstream operations. However, some disadvantages of enzyme catalysis include the ease of deactivation of enzymes in these systems, generally low reaction rates, and low conversions. For example, the immobilized enzymes are easily deactivated in the absence of polar compounds such as water and methanol. Moreover, immobilized enzymes are generally more expensive than chemical catalysts. In response to earlier reports on the low degree of methanolysis using lipases and impossibility to reuse lipase catalysts, Shimada et al.[193, 194] developed a stepwise methanolysis system with immobilized Candida. They suggested that the problem was due to the irreversible inactivation of the lipase upon contact with the concentrated MeOH, which was for the most part insoluble in the reaction mixture. To avoid this deactivation, the first-step methanolysis was conducted at 30 8C in a mixture of oil with only 1/3 molar equivalents of the required amount of MeOH and 4 wt % immobilized lipase by weight of the reaction mixture. The conversion of the oil reached 33.1 % at 7 h. The addition of a second 1/3 molar equivalent of MeOH at 10 h resulted in 66.4 % oil conversion. A third 1/3 molar equivalent of MeOH was added again after a total of 24 h of reaction time, and the reaction was continued for other 24 h. After 48 h, the conversion reached 97.3 %. The methanolysis of cottonseed oil was studied by Royon et al.[195] using the immobilized Candida antarctica lipase (Novozym 435) as catalyst. The authors, using a batch system, were able to significantly increase the reaction rate and yield of ester by adding tert-butyl alcohol to the reaction medium. In fact, the enzyme inhibition caused by the bulk methanol phase was eliminated. Remarkably, a methanolysis yield of 97 % was observed after 24 h at 50 8C with a reaction mixture containing 32.5 % tert-butyl alcohol, 13.5 % methanol, 54 % oil, and 0.017 g enzyme per gram of oil. Ting et al.[196] studied an enzyme/acid-catalyzed hybrid process using soybean oil as feedstock. They adopted a two-stage process consisting of an initial enzymatic hydrolysis followed by an acid-catalyzed esterification. In the first step, 88 % of the substrate was hydrolyzed to FFAs after 5 h using a binary immobilized lipase under optimized conditions (oil/buffer ratio of 2/1 w/w, pH 8, 40 8C). Successive esterification of the FFAs with methanol (FFA/MeOH molar ratio of 1:15) in the presence of 2.5 % sulfuric acid yielded a biodiesel conversion of 99 % after 12 h. Recently, Dizge and Keskinler[197] investigated the effects of pH, molar ratio, and temperature on the enzymatic production of biodiesel from canola oil (0.3 % FFA content and an acid value of 0.6 mgKOH g1) using an immobilized lipase. The yield
ChemSusChem 2009, 2, 278 300

of methyl ester increased with increasing oil/methanol molar ratio until the highest yield (85.8 %) was obtained at a molar ratio of 1:6. A further increase in molar ratio led to a considerable decrease in the yield. The effect of temperature was studied over the range 3070 8C. It was found that 40 8C was the optimal temperature for the reaction, but the activity of the enzyme was dramatically reduced at temperatures above 50 8C. Noureddini et al.[198] studied the enzymatic transesterification of soybean oil with methanol and ethanol. Among nine lipases tested, lipase PS from Pseudomonas cepacia resulted in the highest yield of alkyl esters. Lipase from P. cepacia was further investigated in an immobilized form within a chemically inert, hydrophobic sol-gel support. The gel-entrapped lipase was prepared by polycondensation of hydrolyzed tetramethoxysilane and iso-butyltrimethoxysilane. The optimal conditions for processing 10 g of soybean oil were a temperature of 35 8C, oil/methanol molar ratio of 1:7.5, 0.5 g water, and 475 mg lipase for the reactions with methanol, and a temperature of 35 8C, oil/ethanol molar ratio of 1:15.2, 0.3 g water, and 475 mg lipase for the reaction with ethanol. Under the optimized conditions, methyl and ethyl esters were obtained in 67 and 65 % yield, respectively, after 1 h of reaction using the immobilized lipase. Triglycerides reached negligible levels after the first 30 min of reaction. The immobilized lipase was consistently more active than the free enzyme and proved to be stable when subjected to repeated uses. 3.5. Transesterification in Alternative Solvent Media Most catalytic (homogeneous, heterogeneous, or enzymatic) transesterification systems have to deal with the transfer of reagents through the liquidliquid interface or to the solid catalyst surface. The related mass-transfer limitations usually significantly reduce the rate and efficiency of the reaction. Several methods are known in organic chemistry to facilitate reactions constrained by mass transfer or to simplify the separation of the reagents, products, and catalysts, namely phase-transfer catalysis, micellar catalysis, or the use of alternative solvents (ionic liquids, supercritical fluids, etc.). Whilst the known saponification phenomenon was claimed to be necessary to some extent owing to a co-catalytic micellar effect on the reaction (increasing liquidliquid interface),[199] its presence is usually considered undesirable. On the other hand, the phase-transfer catalysis method, which uses quaternary ammonium salts or crown ethers able to form ion pairs with otherwise coordinated reacting ionic species, although widely applied in biphasic systems, has received scarce attention in transesterification processes[200, 201] and even less for the transesterification of vegetable oils.[202] On the other hand, alternative solvents such as supercritical methanol (350 8C and 3040 MPa) have received much interest in application to the methanolysis of vegetable oils.[203] One of the advantages of supercritical fluids is their high diffusivities, comparable to that of gases, whilst their density is comparable to that of liquids. In addition, no phase separation occurs in the transesterification performed in supercritical methanol,

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

293

S. Miertus et al. which renders this system an ideal solution to overcome masstransfer limitations. In supercritical methanol, biodiesel can be produced even in the absence of a catalyst.[137, 204216] The supercritical methanol process is therefore fast, efficient, and simple in operation allowing over 98 % yields of FAMEs in just a few minutes with no need for further catalyst separation. Other advantages of the supercritical methanol system include its good performance in the presence of water and, in particular, FFAs. The latter, which are initially present (e.g. in waste cooking oils) or formed in situ through hydrolysis of the oil, undergo efficient esterification to biodiesel in the same system. Supercritical fluids can also be excellent reaction media for transesterification in the presence of catalysts. For example, Varma et al.[217, 218] synthesized biodiesel from castor oil with methanol and ethanol using Novozym 435 in supercritical carbon dioxide. Demirbas et al.[219] showed that magnesium oxide can enhance the transesterification reaction of sunflower oil in supercritical methanol. Other alternative solvent systems such as ionic liquids can be useful in the transesterification of vegetable oils so as to facilitate the separation of the catalyst from the reagents and products. Ionic liquids, whose properties can be tuned so that they are immiscible with either aqueous, alcohol, or organic phases, are also known for being efficient supports for organometallic catalysts. An example of such supported ionic liquid catalysis was reported for the methanolysis of soybean oil[220] using Sn(3-hydroxy-2-methyl-4-pyron)2ACHTUNGRE(H2O)2 anchored to 1butyl-3-methylimidazolium hexafluorophosphate. However, this prepared system did not maintain activity (over 90 % activity loss after repeated use) owing to leaching of the catalyst from the ionic-liquid phase. teOil) but requires large-scale production to be competitive with other biodiesels, which means high feedstock availability. A general mechanism for cracking processes that occur during the thermal processing of triglycerides is given in Scheme 4.[222] The first step involves the formation of free fatty

Scheme 4. The mechanism of thermal decomposition of triglycerides. Adapted from Ref. [222].

4. Cracking and Hydroprocessing of Vegetable Oils


Generally, pyrolysis or thermal cracking of organic matter, for example, petroleum fractions, involves the chemical decomposition of organic molecules induced by heat in the absence of air or oxygen to yield smaller molecules. Cracking refers to pyrolytic treatment in the presence of a catalyst, which directs the process mainly towards lower-molecular-weight aliphatic and aromatic hydrocarbons with lower oxygen content. In the case of cracking in the presence of hydrogen and an appropriate catalyst, the process is generally called hydroprocessing. In principle, the well-known cracking and hydroprocessing technologies used in conventional petroleum refineries[221] could be used to treat feedstocks such as pyrolysis or vegetable oils. The interest in cracking and hydroprocessing of vegetable oils to produce fuels has increased recently as a result of a better atom-efficiency of the process and its feedstock tolerance. In fact, the process converts all the organic content into the target product notwithstanding the presence of FFAs and does not produce large quantities of by-products such as glycerol. The hydroprocessing of oils can be cost effective (in fact NExBTL is a commercial hydrotreated vegetable oil by Nes-

acids, which then undergo a series of radical bond-cleavage reactions, including processes of hydrogen transfer, coupling, decarboxylation, and radical termination. The final products range from alkanes, olefins, and aromatics to unsaturated and saturated methyl esters, and gases. The product yield and composition depend on many factors, such as the nature and the quality of the feedstock oil, the operating conditions (type of reactor, temperature, regime, etc.), and the type and the amount of catalyst. Thermal cracking of vegetable oils is generally carried out at temperatures of 350500 8C or higher and at moderate pressures (atmospheric to several bar) to yield around 70 % hydrocarbons. As a rule, an increase in pyrolysis temperature leads to an increase in the selectivity towards the formation of light hydrocarbons. Usually, thermally cracked vegetable oils show better fuel properties than the original oils, however, these are still inferior to that of conventional diesel (Table 3) which suggests that further refining of cracked oils is required. The first application of pyrolysis of vegetable oils refers to an attempt to synthesize petroleum from vegetable oil. Since World War I, many investigators have studied the pyrolysis (thermal cracking) of vegetable oils to obtain products suitable for fuel.[224, 225] In the study by Chang and Wan (1942),[226] tung oil was first saponified with lime and then thermally cracked to yield crude bio-oil, which was refined to diesel fuel and small amounts of gasoline and kerosene. Since then, a number of studies have been dedicated to the thermal or catalytic cracking of soybean,[222, 227] palm,[228231] rapeseed,[232] canola,[229] and other vegetable oils.[203, 228, 233] Many studies have focused on improving the performance of standard catalytic cracking procedures for application to
ChemSusChem 2009, 2, 278 300

294

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production high aromatization rate of the fatty acids caused the formation Diesel fuel of large amounts of aromatics (up to 3040 wt %) in the gaso[223] [222]b line fraction. 51.0 40.0 Tian et al.[230] used a laborato45.6 45.5 ry-scale two-stage riser fluid catalytic cracking unit at 500520 8C 6.7 6.7 using USY and ZSM-5 zeolitemax max 2.82 1.94.1 based catalysts to treat palm and other oils alone or co-fed with vacuum gas oil. For the cracking of palm oil, a conversion of over 97 % was observed, with liquefied petroleum gas (LPG) obtained in 45 % yield, propylene in 23 % yield, and 77.6 % total liquid yield. Co-feeding palm oil with vacuum gas oil gave LPG in 39.1 % yield and propylene in 18.1 % yield, with a total liquid yield of 79.2 %. The oxygen content was very low (about 0.5 %) in the liquid products. At the same time, a high concentration of aromatics (C7C9 predominantly) in the gasoline fraction was observed. Sadramelli and Green[242] used an analytical semi-empirical model to analyze a large number of experimental data for the catalytic cracking of canola oil with HZSM-5, H-mordenite, H-Y, silicalite, aluminium-pillared clay, and silica-alumina catalysts for the production of renewable aromatic hydrocarbons. The results indicated that HZSM-5 catalyst was the most active system. Similar to conventional petroleum oil refining, the fuel properties of the vegetable oil cracking products can be improved if the process is carried out in the presence of hydrogen or followed by treatment with hydrogen. Hydroprocessing is well known in the petrochemical industry and includes catalytic hydrocracking technology and more diffused hydrotreating technology. Whilst the catalyst formulations are similar or the same for both processes, the former promotes cracking and isomerization reactions to a bigger extent as a result of harsher conditions and prolonged residence times, whereas the latter brings about a more significant hydrogenation and hydrogenolysis component leading to the saturation and removal of heteroatoms, such as nitrogen, sulfur, and oxygen. These processes can be sometimes called deoxygenation when applied to oxygenated hydrocarbons, such as vegetable oils or FFAs, as they lead to the removal of oxygen by hydrogenation, decarboxylation, and decarbonylation. Therefore, fully deoxygenated highcetane hydrocarbons suitable for use in diesel engines are formed. As typical hydrocracking/hydrotreating catalysts combine cracking and hydrogenation components, metal catalysts supported on various acidic oxides are normally used such as Ni-Mo, Co-Mo, as well as Pd and other metals on silica and alumina. Similar to the cracking pathways outlined in Scheme 4, the initial stage of hydrocracking of triglycerides would be saturation of fatty acid chains and rapture of the ester groups leading to the formation of FFAs (see Scheme 5). In principle, the latter can be further cracked, but at lower temperatures these rather undergo deoxygenation and hydrogenation (hydrotreat-

Table 3. Fuel properties of thermally cracked soybean oil from two different studies. Property Soybean oil [223] Cetane number Higher heating value [MJ kg1] Pour point [8C] Viscosity [cSt] (37.8 8C) 38.0 39.3 12.2 32.6 [222] 37.9 39.6 12.2 32.6 Cracked soybean oil [223] [222] 43.0 40.6 4.4 7.74 43.0 40.3 7.2 10.2

vegetable oils using standard or modified cracking catalysts, such as those based on activated alumina,[234] aluminosilicates,[68, 235237] zeolites,[238240] and pillared clays.[229] Ooi et al.[68, 235237] investigated pure and composite HZSM-5 and MCM-41/SBA-15 type zeolites in the catalytic conversion of a mixture of palm-oil-based fatty acids to a liquid fuel. The reaction was conducted in a fixed-bed microreactor at a temperature of 450 8C and at a weight hourly space velocity of 2.5 h1. The maximum conversion of the fatty acids mixture was 98 wt % over a composite catalyst comprising 30 wt % coating of purely siliceous SBA-15 over HZSM-5. The maximum gasoline fraction yield of 44 wt % was obtained with a HZSM-5based composite catalyst coated with 20 wt % pure siliceous MCM-41. The selectivity for the production of benzene, toluene, and xylene (BTX) aromatics in the organic liquid product was enhanced over the mesophase aluminum-containing composite catalysts relative to the mesoporous MCM-41 or microporous HZSM-5 catalysts alone. Furthermore, Al-doped SBA-15 is of potential interest as its thermal stability is significantly higher than that of Al-MCM-41 but with similar cracking activity. Katikaneni et al.[239] studied the production of hydrocarbons both from wood-derived and vegetable oils using aluminophosphate as a catalyst. The results indicated that the optimum hydrocarbon yields were low as compared to those obtained with HZSM-5. However, aluminophosphate showed an interesting selectivity towards the production of aromatic compounds, such as benzene (with wood-derived oil) and toluene and xylenes (with canola oil), as well as towards important gaseous hydrocarbons, such as ethylene, propylene, and n-butane. The co-feeding of vegetable oils with fossil feedstocks to fluid catalytic cracking units proved to be a simple and cheap route, even though dedicated units could leave more flexibility and the possibility to improve performances and yields. However, the experimental conditions used in most laboratoryscale studies are far from those of the fluid catalytic cracking units where the same catalyst can show significantly lower performances as a result of thermodynamic limitations. The cracking of rapeseed oil under realistic fluid catalytic cracking conditions and using a commercial equilibrium catalyst was investigated, for instance, recently by Dupain et al.[241] The data indicated that the triglycerides were predominately converted into fatty acids through radical cracking reactions within 50 ms at 485585 8C. Relatively high amounts of coke were formed. The
ChemSusChem 2009, 2, 278 300

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

295

S. Miertus et al.

Scheme 5. General pathways of hydrotreatment of vegetable oils. Adapted from Ref. [243].

ing) to produce suitable diesel-like hydrocarbons (C15C18). Therefore, the process can employ less harsh conditions than conventional cracking over zeolites. For example, Huber et al.[243] hydrotreated vegetable oils with hydrogen at 300 450 8C and 50 bar pressure over sulfided Ni-Mo supported on alumina catalysts. They noted the presence of hydrogenation, decarboxylation, decarbonylation, and some cracking pathways which took place simultaneously to afford straight-chain C15 C18 alkanes in 71 % yield (against the 95 % maximum theoretical yield). Sebos et al.[244] used Co-Mo supported on alumina at 300345 8C and 30 bar to produce biodiesel from refined cottonseed oil. On the other hand, hydrotreatment normally involves higher pressures than those used in catalytic cracking (up to 250 bar).[245] The possibility of co-feeding vegetable oils together with heavy vacuum oil or vacuum gas oil have been evaluated in some studies.[246] It was also indicated that such a process is normally less efficient than treatment of vacuum gas oil or vegetable oil alone as a result of the difference in optimal conditions.[247] In the case of FFAs, their deoxygenation can be based mainly on decarboxylation and therefore can be carried out under milder conditions and in the absence of hydrogen. For example, the conversion of C15C18 FFAs to the corresponding straight-chain hydrocarbons suitable for use as fuel over Pd supported on activated carbons at moderate temperatures and pressures (around 300 8C, 617 bar) has been reported.[248] Like in the case of cracking, the hydrocracking of vegetable oils can be also carried out in conventional hydrotreating units using commercial catalysts. For example, Stumborg et al.[249] reported that the middle distillate of the cracking product with cetane number 5590 % and sulfur content less than 10 ppm was obtained in yields of 7080 % using a proprietary Canadian technology.

used sodium hydroxide and sodium methoxide offer reasonably short reaction times and good conversions and product yields. However, the effectiveness of homogeneous base catalysts in transesterification is limited to feedstocks that contain no water and free fatty acids. The combination of acid-catalyzed esterification with base-catalyzed transesterification could be a solution for feedstocks with a high FFA content, as well as coupling hydrolysis with esterification processes; however, a more complicated processing is required. As most systems using homogeneous catalysts require batch-mode processing and a number of downstream processes to remove residues of catalysts and to purify products, the overall process can be still improved from both environmental and economic points of view. For example, a higher purity of glycerol and its valorization are typical issues. Homogenous catalysts can be replaced by heterogeneous catalysts in order to run the transesterification process in continuous mode and avoid unfavorable saponification phenomena and related product-separation issues, as well as to allow other simple downstream processing steps and recovery of the catalyst. Overall, heterogeneous catalysts are more environmentally friendly than homogeneous ones and represent a

5. Conclusion and Outlook


Biodiesel production technologies based on homogenous catalysts are simple in operation and cost effective. Commonly

more sustainable mode of resource management. However, there are still improvements to be made on the formulations as to allow their higher intrinsic efficiency in this reaction. Generally, heterogeneous catalysts are less active and suffer from deactivation phenomena. Nevertheless, as a result of an unlimited number of possible materials with a flexibility of properties to be used as catalysts, massive research afforts are currently underway to develop compositions of materials with better acidic or basic properties, improved pore structure, better lipophilic properties, and more stable performance or tolerance to the reaction conditions and components. Innovative materials such as zeolites, heteropolyacids, hydrotalcites, and mesostructured mixed oxides, and their compositions with various basic or acidic species of both organic and inorganic nature, are being explored for transesterification and esterificaChemSusChem 2009, 2, 278 300

296

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Catalytic Applications in Biodiesel Production tion processes. Several industrial biodiesel manufacturers have already adopted heterogeneous processes proving that a considerable amount of progress was made in this direction. For instance, AxensIFP Group Technologies commercialized Esterfip-H, an innovative heterogeneously catalyzed technology for the production of high-quality biodiesel that also allows the production of good-quality glycerol. Two Esterfip-H plants in France and Sweden have been built so far, and several other plants are in preparation.[250] NOVA Biosource Fuels offers a patented, heterogeneous catalytic conversion process to treat both free fatty acids and triglycerides.[251] Enzymes can be called new-generation catalysts for the production of first-generation biodiesel from vegetable oils, as they offer a further set of improvements to the traditional process. Enzymes such as lipases promote the energetically favorable low-temperature transesterification process with yet lower environmental impact. As in the case of heterogeneous catalysts, the saponification and related separation issues are avoided, but, additionally, both the transesterification of triglycerides and esterification of FFAs can be run in one pot and at a lower methanol to oil ratio. This eliminates the need of feedstock pretreatment and affords biodiesel and glycerol of better purity. However, generally, the reaction rates and conversions are lower than those observed with traditional homogeneous or heterogeneous methods. Research is currently directed towards improvements in the lifetime of enzymes and their resistance to deactivation (e.g. with methanol), as well as towards the possibility to immobilize them for applications in continuous processing. As these developments are still at an early stage and even unstable enzyme catalysts are very costly, enzyme transesterification technology is still far from commercial full-scale operation. A good use of the existing progress in heterogeneous catalysis is represented by cracking and hydrocracking technologies, which can be used to process vegetable oils into deoxygenated hydrocarbon fuels. The general advantages of the cracking technology, and specifically in application to vegetable oils, is that it can be run in continuous mode and is relatively independent of the composition of the feedstock. In the case of vegetable oils, the latter means the possibility to make equal use of the whole feedstock mass (triglycerides and FFAs are processed). Therefore, complex steps of liquid product separation and catalyst recovery are eliminated and there is no need to valorize the by-product glycerol. Even though the selectivity of the process in the biodiesel fraction is lower than 80 %, other reaction products can be integrated in the existing petroleum refinery. The latter possibility is probably the main advantage of the process, even though it is quite energy-consuming and its suitability to the vegetable feedstock needs improvements. Given ample feedstock availability and slightly modified catalyst formulations and reactor design to address more acidic feedstocks, the process can be cost-competitive. In fact, industrial-scale facilities are now in operation and a huge interest in related research indicates a promising future for the hydrocracking technology.

Acknowledgements
A.S., F.K., and P.F. gratefully acknowledge support through the fellowship and research programs provided by ICS-UNIDO. Keywords: biofuels enzyme catalysis heterogeneous catalysis homogeneous catalysis transesterification
[1] J. Wassell Jr. , T. P. Dittmer, Energy Policy 2006, 34, 3993. [2] J. Hill, E. Nelson, D. Tilman, S. Polasky, D. Tiffany, Proc. Natl. Acad. Sci. USA 2006, 103, 11206. [3] A. Demirbas, Energy Explor. Exploit. 2003, 21, 475. [4] A. H. Demirbas, I. Demirbas, Energy Convers. Manage. 2007, 48, 2386. [5] A. Demirbas, Energy Policy 2007, 35, 4661. [6] Data from the Energy Information Administration: http://tonto.eia.doe.gov/oog/info/gdu/dieselpump.html. [7] The Future of Liquid Biofuels for APEC Economies, Energy Working Group, APEC, May 2008. [8] J. M. Marchetti, V. U. Miguel, A. F. Errazu, Fuel Process. Technol. 2008, 89, 740. [9] M. J. Haas, A. J. McAloon, W. C. Yee, T. A. Foglia, Bioresour. Technol. 2006, 97, 671. [10] From 1st to 2nd generation biofuel technologies. An Overview of Current Industry and RD&D Activities, OECD/IEA, November 2007: www.iea.org. [11] D. H. Lee, D. J. Lee, Energy Fuels 2008, 22, 177. [12] Data from www.biodieselindustries.com. [13] L. C. Meher, D. Vidya Sagar, S. N. Naik, Renewable Sustainable Energy Rev. 2006, 10, 248. [14] P. T. Vasudevan, M. Briggs, J. Ind. Microbiol. Biotechnol. 2008, 35, 421. [15] G. Knothe, J. Van Gerpen, J. Krahl, The Biodiesel Handbook, AOCS, Champaign, IL, 2005. [16] A. H. Demirbas, Energy Convers. Manage. 2003, 44, 2093. [17] A. Srivastava, R. Prasad, Renewable Sustainable Energy Rev. 2000, 4, 111. [18] B. Freedman, E. H. Pryde, T. L. Mounts, J. Am. Oil Chem. Soc. 1984, 61, 1638. [19] S. Fernando, C. Hall, S. Jha, Energy Fuels 2006, 20, 376. [20] H. Hazar, Renewable Energy 2009, 34, 1533. [21] G. Knothe, C. A. Sharp, T. W. Ryan III, Energy Fuels 2006, 20, 403. [22] M. Nabi, M. Akhter, M. M. Shahadat, Bioresour. Technol. 2006, 97, 372. [23] C. D. Rakopoulos, K. A. Antonopoulos, D. C. Rakopoulos, D. T. Hountalas, E. G. Giakoumis, Energy Convers. Manage. 2006, 47, 3272. [24] R. O. Dunn, Trans. ASABE 2006, 49, 1633. [25] G. Knothe, Fuel Process. Technol. 2007, 88, 669. [26] H. Noureddini, D. Zhu, J. Am. Oil Chem. Soc. 1997, 74, 1457. [27] M. Mittelbach, B. Trathnigg, Fat Sci. Technol.. 1990, 92, 145. [28] K. Komers, F. Skopal, R. Stloukal, J. Machek, Eur. J. Lipid Sci. Technol. 2002, 104, 728. [29] S. K. Karmee, D. Chandna, R. Ravi, A. Chadha, J. Am. Oil Chem. Soc. 2006, 83, 873. [30] D. Darnoko, M. Cheryan, J. Am. Oil Chem. Soc. 2000, 77, 1263. [31] T. Leevijit, W. Wisutmmethangoon, G. Prateepchaikul, C. Tongurai, M. Allen, The Joint International Conference on Sustainable Energy and Environment (SEE), December 13, 2004, Hua Hin, Thailand, 277 281. [32] D. G. B. Boocock, S. K. Konar, V. Mao, C. Lee, S. Buligan, J. Am. Oil Chem. Soc. 1998, 75, 1167. [33] E. Bikou, A. Louloudi, N. Papayannakos, Chem. Eng. Technol. 1999, 22, 70. [34] A. Cornish-Bowden, Principles of Enzyme Kinetics, Butterworths, London, 1976. [35] S. Al-Zuhair, F. W. Ling, L. S. Jun, Process Biochem. 2007, 42, 951. [36] Y. Xu, W. Du, D. Liu, J. Mol. Catal. B 2005, 32, 241. [37] P. Mensah, J. L. Gainer, G. Carta, Biotechnol. Bioeng. 1998, 60, 434. [38] V. Dossat, D. Combes, A. Marty, Enzyme Microb. Technol. 1999, 25, 194. [39] A. Zaks, A. M. Klibanov, Science 1984, 224, 1249. [40] S. Gryglewicz, Bioresour. Technol. 1999, 70, 249. [41] E. Leclercq, A. Finiels, C. Moreau, J. Am. Oil Chem. Soc. 2001, 78, 1161. [42] X. Yuan, J. Liu, G. Zeng, J. Shi, J. Tong, G. Huang, Renewable Energy 2008, 33, 1678.

ChemSusChem 2009, 2, 278 300

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

297

S. Miertus et al.
[43] F. R. Abreu, D. G. Lima, E. H. Hamu, S. Einloft, J. C. Rubim, P. A. Z. Suarez, J. Am. Oil Chem. Soc. 2003, 80, 601. [44] B. Freedman, R. O. Butterfield, E. H. Pryde, J. Am. Oil Chem. Soc. 1986, 63, 1375. [45] S. Furuta, H. Matsuhashi, K. Arata, Biomass Bioenergy. 2006, 30, 870. [46] L. Guerreiro, J. E. Castanheiro, I. M. Fonseca, R. M. Martin-Aranda, A. M. Ramos, J. Vital, Catal. Today 2006, 118, 166. [47] H. Li, W. Xie, J. Am. Oil Chem. Soc. 2008, 85, 655. [48] X. Liu, X. Piao, Y. Wang, S. Zhu, Energy Fuels 2008, 22, 1313. [49] X. Liu, X. Piao, Y. Wang, S. Zhu, H. He, Fuel 2008, 87, 1076. [50] X. Liu, H. He, Y. Wang, S. Zhu, X. Piao, Fuel 2008, 87, 216. [51] C. V. McNeff, L. C. McNeff, B. Yan, D. T. Nowlan, M. Rasmussen, A. E. Gyberg, B. J. Krohn, R. L. Fedie, T. R. Hoye, Appl. Catal. A: Gen. 2008, 343, 39. [52] U. Schuchardt, R. M. Vargas, G. Gelbard, J. Mol. Catal. A: Chem. 1996, 109, 37. [53] G. J. Suppes, M. A. Dasari, E. J. Doskocil, P. J. Mankidy, M. J. Goff, Appl. Catal. A: Gen. 2004, 257, 213. [54] W. Xie, H. Li, J. Mol. Catal. A: Chem. 2006, 255, 1. [55] W. Xie, Z. Yang, Catal. Lett. 2007, 117, 159. [56] W. Xie, H. Peng, L. Chen, Appl. Catal. A: Gen. 2006, 300, 67. [57] Z. Yang, W. Xie, Fuel Process. Technol. 2007, 88, 631. [58] X. Bo, X. Guomin, C. Lingfeng, W. Ruiping, G. Lijing, Energy Fuels 2007, 21, 3109. [59] P. Cao, M. A. Dube, A. Y. Tremblay, Biomass Bioenergy 2008, 32, 1028. [60] T. L. Chew, S. Bhatia, Bioresour. Technol. 2008, 99, 7911. [61] L. Gao, B. Xu, G. Xiao, J. Lv, Energy Fuels 2008, 22, 3531. [62] S. F. A. Halim, A. H. Kamaruddin, W. J. N. Fernando, Bioresour. Technol. 2009, 100, 710. [63] J. Kansedo, K. T. Lee, S. Bhatia, Biomass Bioenerg. 2009, 33, 271.. [64] J. Kansedo, K. T. Lee, S. Bhatia, Chem. Eng. Technol. 2008, 31, 993. [65] X. Lu, D. Bao, M. Su, Pet. Process. Petrochem. 2008, 39, 55. [66] K. Noiroj, P. Intarapong, A. Luengnaruemitchai, S. Jai-In, Renewable Energy 2009, 34, 1145. [67] Y. S. Ooi, R. Zakaria, A. R. Mohamed, S. Bhatia, Energy Fuels 2004, 18, 1555. [68] Y. S. Ooi, R. Zakaria, A. R. Mohamed, S. Bhatia, Appl. Catal. A: Gen. 2004, 274, 15. [69] G. Prateepchaikul, M. L. Allen, T. Leevijit, K. Thaveesinsopha, Songklanakarin J. Sci. Technol. 2007, 29, 1551. [70] C. W. Puah, Y. M. Choo, A. N. Ma, C. H. Chuah, Lipids 2006, 41, 305. [71] C. Tongurai, K. Chuntorng-Orn, Songklanakarin J. Sci. Technol. 2007, 29, 1583. [72] S. C. M. Dos Reis, E. R. Lachter, R. S. V. Nascimento, J. Rodrigues, M. G. Reid, J. Am. Oil Chem. Soc. 2005, 82, 661. [73] S. Benjapornkulaphong, C. Ngamcharussrivichai, K. Bunyakiat, Chem. Eng. J. 2009, 145, 468. [74] K. J. Harrington, C. Arcy-Evans, Ind. Eng. Chem. Prod. Res. Dev. 1985, 24, 314. [75] M. J. Ramos, A. Casas, L. Rodriguez, R. Romero, A. Perez, Appl. Catal. A: Gen. 2008, 346, 79. [76] M. C. G. Albuquerque, J. Santamaria-Gonzalez, J. M. Merida-Robles, R. Moreno-Tost, E. Rodriguez-Castellon, D. C. S. Azevedo, C. L. Cavalcante, Jr., P. Maireles-Torres, Appl. Catal. A: Gen. 2008, 347, 162. [77] G. Arzamendi, I. Campo, E. Arguinarena, M. Sanchez, M. Montes, L. M. Gandia, J. Chem. Technol. Biotechnol. 2008, 83, 862. [78] K. G. Georgogianni, M. G. Kontominas, P. J. Pomonis, D. Avlonitis, V. Gergis, Fuel Process. Technol. 2008, 89, 503. [79] D. Frascari, M. Zuccaro, D. Pinelli, A. Paglianti, Energy Fuels 2008, 22, 1493. [80] U. Rashid, F. Anwar, B. R. Moser, S. Ashraf, Biomass Bioenergy 2008, 32, 1202. [81] M. Verziu, B. Cojocaru, J. Hu, R. Richards, C. Ciuculescu, P. Filip, V. I. Parvulescu, Green Chem. 2008, 10, 373. [82] G. Arzamendi, E. Arguinarena, I. Campo, S. Zabala, L. M. Gandia, Catal. Today 2008, 133135, 305. [83] G. Sunita, B. M. Devassy, A. Vinu, D. P. Sawant, V. V. Balasubramanian, S. B. Halligudi, Catal. Commun. 2008, 9, 696. [84] M. C. G. Albuquerque, I. Jimenez-Urbistondo, J. Santamaria-Gonzalez, J. M. Merida-Robles, R. Moreno-Tost, E. Rodriguez-Castellon, A. Jimenez-Lopez, D. C. S. Azevedo, C. L. Cavalcante, Jr., P. Maireles-Torres, Appl. Catal. A: Gen. 2008, 334, 35. D. I. Jordanov, Y. K. Dimitrov, P. S. T. Petkov, S. K. Ivanov, Oxid. Commun. 2007, 30, 300. G. Vicente, M. Martinez, J. Aracil, Bioresour. Technol. 2007, 98, 1724. M. L. Granados, M. D. Z. Poves, D. M. Alonso, R. Mariscal, F. C. Galisteo, R. Moreno-Tost, J. Santamaria, J. L. G. Fierro, Appl. Catal. B 2007, 73, 317. R. Wang, S. Yang, S. Yin, B. Song, P. S. Bhadury, W. Xue, S. Tao, Z. Jia, D. Liu, L. Gao, Front. Chem. Eng. China 2008, 2, 468. N. C. Om Tapanes, D. A. Gomes Aranda, J. W. de Mesquita Carneiro, O. A. Ceva Antunes, Fuel 2008, 87, 2286. H. J. Berchmans, S. Hirata, Bioresour. Technol. 2008, 99, 1716. F. R. Abreu, D. G. Lima, E. H. Hamu, C. Wolf, P. A. Z. Suarez, J. Mol. Catal. A: Chem. 2004, 209, 29. G. Steinke, S. Schonwiese, K. D. Mukherjee, J. Am. Oil Chem. Soc. 2000, 77, 367. J. M. Encinar, J. F. Gonzalez, J. J. Rodriguez, A. Tejedor, Energy Fuels 2002, 16, 443. J. M. Encinar, J. F. Gonzalez, E. Sabio, M. J. Ramiro, Ind. Eng. Chem. Res. 1999, 38, 2927. T. Eevera, K. Rajendran, S. Saradha, Renewable Energy 2009, 34, 762. L. C. Meher, M. G. Kulkarni, A. K. Dalai, S. N. Naik, Eur. J. Lipid Sci. Technol. 2006, 108, 389. Q. Li, W. Du, D. Liu, Appl. Microbiol. Biotechnol. 2008, 80, 749. A. B. M. S. Hossain, A. Salleh, A. N. Boyce, P. Chowdhury, M. Naqiuddin, Am. J. Biochem. Biotechnol. 2008, 4, 250. A. Demirbas, Energy Sources Part A 2009, 31, 163. P. T. Vasudevan, M. Briggs, J. Ind. Microbiol. Biotechnol. 2008, 35, 421. H. Xu, X. Miao, Q. Wu, J. Biotechnol. 2006, 126, 499. X. Miao, Q. Wu, Bioresour. Technol. 2006, 97, 841. L. Poisson, F. Ergan, J. Biotechnol. 2001, 91, 75. E. H. Belarbi, E. Molina, Y. Chisti, Process Biochem. 2000, 35, 951. R. Altin, S. Cetinkaya, H. S. Yucesu, Energy Convers. Manage. 2001, 42, 529. S. Zullaikah, C. C. Lai, S. R. Vali, Y. H. Ju, Bioresour. Technol. 2005, 96, 1889. O. E. Ikwuagwu, I. C. Ononogbu, O. U. Njoku, Ind. Crops Prod. 2000, 12, 57. A. S. Ramadhas, S. Jayaraj, C. Muraleedharan, Fuel 2005, 84, 335. A. Saydut, M. Z. Duz, C. Kaya, A. B. Kafadar, C. Hamamci, Bioresour. Technol. 2008, 99, 6656. N. Usta, Biomass Bioenergy 2005, 28, 77. O. J. Alamu, M. A. Waheed, S. O. Jekayinfa, Fuel 2008, 87, 1529. C. Ngamcharussrivichai, P. Totarat, K. Bunyakiat, Appl. Catal. A: Gen. 2008, 341, 77. C. Ngamcharussrivichai, W. Wiwatnimit, S. Wangnoi, J. Mol. Catal. A: Chem. 2007, 276, 24. R. E. Armenta, M. Vinatoru, A. M. Burja, J. A. Kralovec, C. J. Barrow, J. Am. Oil Chem. Soc. 2007, 84, 1045. M. E. D. S. Padilha, W. Augusto-Ruiz, Cienc. Tecnol. Aliment. 2007, 27, 285. Y. C. Sharma, B. Singh, S. N. Upadhyay, Fuel 2008, 87, 2355. A. Demirbas, Energy Convers. Manage. 2008, 49, 2106. C. Devendra, Outlook Agric. 2004, 33, 157. M. Canakci, J. Van Gerpen, Trans. Am. Soc. Agr. Eng. 2001, 44, 1429. G. Knothe, K. R. Steidley, Fuel 2005, 84, 1059. G. Knothe, Fuel Process. Technol. 2005, 86, 1059. M. K. Smith, US 2,486,444, 1949. A. A. Refaat, N. K. Attia, H. A. Sibak, S. T. El Sheltawy, G. I. ElDiwani, Int. J. Environ. Sci. Technol. 2008, 5, 75. H. Noureddini, D. Harkey, V. Medikonduru, J. Am. Oil Chem. Soc. 1998, 75, 1775. A. P. Singh, B. B. He, J. C. Thompson, J. H. Van Gerpen, Appl. Eng. Agr. 2006, 22, 597. M. P. Dorado, E. Ballesteros, J. A. DeAlmeida, C. Schellert, H. P. Lohrlein, R. Krause, Trans. Am. Soc. Agr. Eng. 2002, 45, 525. F. Ma, L. D. Clements, M. A. Hanna, Bioresour. Technol. 1999, 69, 289. F. Ma, L. D. Clements, M. A. Hanna, Trans. Am. Soc. Agr. Eng. 1998, 41, 1261. J. Qian, F. Wang, S. Liu, Z. Yun, Bioresour. Technol. 2008, 99, 9009.

[85] [86] [87]

[88] [89] [90] [91] [92] [93] [94] [95] [96] [97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] [113] [114] [115] [116] [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] [127] [128] [129]

298

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 278 300

Catalytic Applications in Biodiesel Production


[130] G. Vicente, M. Martinez, J. Aracil, Bioresour. Technol. 2004, 92, 297. [131] L. Bournay, D. Casanave, B. Delfort, G. Hillion, J. A. Chodorge, Catal. Today 2005, 106, 190. [132] H. J. Kim, B. S. Kang, M. J. Kim, Y. M. Park, D. K. Kim, J. S. Lee, K. Y. Lee, Catal. Today 2004, 9395, 315. [133] D. E. Lopez, J. Goodwin, D. A. Bruce, E. Lotero, Appl. Catal. A: Gen. 2005, 295, 97. [134] A. A. Kiss, F. Omota, A. C. Dimian, G. Rothenberg, Top. Catal. 2006, 40, 141. [135] A. A. Kiss, A. C. Dimian, G. Rothenberg, Energy Fuels 2008, 22, 598. [136] Y. Watanabe, Y. Shimada, A. Sugihara, H. Noda, H. Fukuda, Y. Tominaga, J. Am. Oil Chem. Soc. 2000, 77, 355. [137] H. Fukuda, A. Kondo, H. Noda, J. Biosci. Bioeng. 2001, 92, 405. [138] K. Jacobson, R. Gopinath, L. C. Meher, A. K. Dalai, Appl. Catal. B 2008, 85, 86. [139] L. L. Lima, J. d. A. Goncalez, A. P. G. Encarnacao, F. Antoniosi, R. Nelson, D. A. G. Aranda, XVII Congresso Brasileiro de Engenharia Quimica (Recife, Brazil), September 1417, 2008. [140] M. J. Haas, K. M. Scott, J. Am. Oil Chem. Soc. 2007, 84, 197. [141] N. E. Leadbeater, L. M. Stencel, Energy Fuels 2006, 20, 2281. [142] N. E. Leadbeater, T. M. Barnard, L. M. Stencel, Energy Fuels 2008, 22, 2005. [143] J. Hernando, P. Leton, M. P. Matia, J. L. Novella, J. Varez-Builla, Fuel 2007, 86, 1641. [144] N. Azcan, A. Danisman, Fuel 2008, 87, 1781. [145] Z. Tang, L. Wang, J. Yang, Eur. J. Lipid Sci. Technol. 2008, 110, 747. [146] U. Schuchardt, R. M. Vargas, G. Gelbard, J. Mol. Catal. A 1995, 99, 65. [147] U. Schuchardt, R. Sercheli, R. M. Vargas, J. Braz. Chem. Soc. 1998, 9, 199. [148] A. Cerro, A. Corma, S. Iborra, J. P. Mez, Appl. Catal. A: Gen. 2008, 346, 52. [149] E. Lotero, Y. Liu, D. E. Lopez, K. Suwannakarn, D. A. Bruce, J. Goodwin, Ind. Eng. Chem. Res. 2005, 44, 5353. [150] S. Zheng, M. Kates, M. A. Dube, D. D. McLean, Biomass Bioenergy 2006, 30, 267. [151] D. A. G. Aranda, R. T. P. Santos, N. C. O. Tapanes, A. L. D. Ramos, O. A. C. Antunes, Catal. Lett. 2008, 122, 20. [152] C. Mazzocchia, G. Modica, A. Kaddouri, R. Nannicini, C. R. Chim. 2004, 7, 601. [153] S. L. Barbosa, M. J. Dabdoub, G. R. Hurtado, S. I. Klein, A. C. M. Baroni, C. Cunha, Appl Catal A: Gen. 2006, 313, 146. [154] T. S. Koh, K. H. Chung, J. Kor. Ind. Eng. Chem. 2008, 19, 214. [155] Q. Shu, B. Yang, H. Yuan, S. Qing, G. Zhu, Catal. Commun. 2007, 8, 2159. [156] F. Chai, F. Cao, F. Zhai, Y. Chen, X. Wang, Z. Su, Adv. Synth. Catal. 2007, 349, 1057. [157] L. Xu, Y. Wang, X. Yang, X. Yu, Y. Guo, J. H. Clark, Green Chem. 2008, 10, 746. [158] F. Cao, Y. Chen, F. Zhai, J. Li, J. Wang, X. Wang, S. Wang, W. Zhu, Biotechnol. Bioeng. 2008, 101, 93. [159] A. Alsalme, E. F. Kozhevnikova, I. V. Kozhevnikov, Appl. Catal. A: Gen. 2008, 349, 170. [160] A. A. Kiss, A. C. Dimian, G. Rothenberg, Adv. Synth. Catal. 2006, 348, 75. [161] F. Omota, A. C. Dimian, A. Bliek, Chem. Eng. Sci. 2003, 58, 3175. [162] G. D. Yadav, A. D. Murkute, J. Catal. 2004, 224, 218. [163] K. Suwannakarn, E. Lotero, J. Goodwin, C. Lu, J. Catal. 2008, 255, 279. [164] S. Furuta, H. Matsuhashi, K. Arata, Appl. Catal. A: Gen. 2004, 269, 187. [165] S. Furuta, H. Matsuhashi, K. Arata, Catal. Commun. 2004, 5, 721. [166] J. Jitputti, B. Kitiyanan, P. Rangsunvigit, K. Bunyakiat, L. Attanatho, P. Jenvanitpanjakul, Chem. Eng. J. 2006, 116, 61. [167] Y. Du, S. Liu, Y. Ji, Y. Zhang, S. Wei, F. Liu, F. S. Xiao, Catal. Lett. 2008, 124, 133. [168] M. H. Zong, Z. Q. Duan, W. Y. Lou, T. J. Smith, H. Wu, Green Chem. 2007, 9, 434. [169] W. Y. Lou, M. H. Zong, Z. Q. Duan, Bioresour. Technol. 2008, 99, 8752. [170] N. Shibasaki-Kitakawa, H. Honda, H. Kuribayashi, T. Toda, T. Fukumura, T. Yonemoto, Bioresour. Technol. 2007, 98, 416. [171] I. K. Mbaraka, B. H. Shanks, J. Catal. 2005, 229, 365. [172] I. K. Mbaraka, B. H. Shanks, J. Catal. 2006, 244, 78. [173] I. K. Mbaraka, D. R. Radu, V. S. Y. Lin, B. H. Shanks, J. Catal. 2003, 219, 329. [174] Y. Huang, G. Qi, J. Qian, L. Feng, Acta Polym. Sin. 2001, 213. [175] D. J. Darensbourg, M. J. Adams, J. C. Yarbrough, Inorg. Chem. 2001, 40, 6543. [176] R. Srivastava, D. Srinivas, P. Ratnasamy, J. Catal. 2006, 241, 34. [177] P. S. Sreeprasanth, R. Srivastava, D. Srinivas, P. Ratnasamy, Appl. Catal. A: Gen. 2006, 314, 148. [178] A. Corma, S. Iborra, S. Miquel, J. Primo, J. Catal. 1998, 173, 315. [179] F. Cavani, F. Trifir, A. Vaccari, Catal. Today 1991, 11, 173. [180] J. I. Di Cosimo, V. K. Diez, M. Xu, E. Iglesia, C. R. Apesteguia, J. Catal. 1998, 178, 499. [181] W. Xie, H. Peng, L. Chen, J. Mol. Catal. A: Chem. 2006, 246, 24. [182] D. G. Cantrell, L. J. Gillie, A. F. Lee, K. Wilson, Appl. Catal. A: Gen. 2005, 287, 183. [183] A. Corma, S. B. A. Hamid, S. Iborra, A. Velty, J. Catal. 2005, 234, 340. [184] C. R. V. Reddy, R. Oshel, J. G. Verkade, Energy Fuels 2006, 20, 1310. [185] A. Kawashima, K. Matsubara, K. Honda, Bioresour. Technol. 2009, 100, 696. [186] W. Xie, X. Huang, Catal. Lett. 2006, 107, 53. [187] T. Ebiura, T. Echizen, A. Ishikawa, K. Murai, T. Baba, Appl. Catal. A: Gen. 2005, 283, 111. [188] G. J. Suppes, K. Bockwinkel, S. Lucas, J. B. Botts, M. H. Mason, J. A. Heppert, J. Am. Oil Chem. Soc. 2001, 78, 139. [189] S. K. F. Peter, R. Ganswindt, H. P. Neuner, E. Weidner, Eur. J. Lipid Sci. Technol. 2002, 104, 324. [190] E. Li, V. Rudolph, Energy Fuels 2008, 22, 145. [191] X. Liu, X. Piao, Y. Wang, S. Zhu, Pet. Process. Petrochem. 2007, 38, 21. [192] I. N. Martyanov, A. Sayari, Appl. Catal. A: Gen. 2008, 339, 45. [193] Y. Shimada, Y. Watanabe, A. Sugihara, Y. Tominaga, J. Mol. Catal. B 2002, 17, 133. [194] Y. Watanabe, Y. Shimada, A. Sugihara, Y. Tominaga, J. Am. Oil Chem. Soc. 2001, 78, 703. [195] D. Royon, M. Daz, G. Ellenrieder, S. Locatelli, Bioresour. Technol. 2007, 98, 648. [196] W. J. Ting, C. M. Huang, N. Giridhar, W. T. Wu, J. Chin. Inst. Chem. Eng. 2008, 39, 203. [197] N. Dizge, B. Keskinler, Biomass Bioenergy 2008, 32, 1274. [198] H. Noureddini, X. Gao, R. S. Philkana, Bioresour. Technol. 2005, 96, 769. [199] C. Stavarache, M. Vinatoru, R. Nishimura, Y. Maeda, Ultrason. Sonochem. 2005, 12, 367. [200] M. Ikunaka, Org. Process Res. Dev. 2008, 12, 698. [201] E. Angeletti, P. Tundo, P. Venturello, J. Org. Chem. 1983, 48, 4106. [202] M. E. Halpern, D. Crick, US 6833463, 2008. [203] I. C. P. Fortes, P. J. Baugh, J. Braz. Chem. Soc. 1999, 10, 469. [204] D. Kusdiana, S. Saka, Bioresour. Technol. 2004, 91, 289. [205] J. M. Marchetti, V. U. Miguel, A. F. Errazu, Renewable Sustainable Energy Rev. 2007, 11, 1300. [206] Y. Warabi, D. Kusdiana, S. Saka, Bioresour. Technol. 2004, 91, 283. [207] D. Kusdiana, S. Saka , J. Chem. Eng. Jpn. 2001, 34, 383. [208] W. Cao, H. Han, J. Zhang, Fuel 2005, 84, 347. [209] A. Demirbas, Energy Convers. Manage. 2006, 47, 2271. [210] D. Kusdiana, S. Saka, Appl. Biochem. Biotechnol. 2004, 115, 781. [211] H. He, T. Wang, S. Zhu, Fuel 2007, 86, 442. [212] K. Bunyakiat, S. Makmee, R. Sawangkeaw, S. Ngamprasertsith, Energy Fuels 2006, 20, 812. [213] Y. Warabi, D. Kusdiana, S. Saka, Appl. Biochem. Biotechnol. 2004, 113 116, 793. [214] D. Kusdiana, S. Saka, Fuel 2001, 80, 693. [215] D. Kusdiana, S. Saka, J. Chem. Eng. Jpn. 2001, 34, 383. [216] S. Saka, D. Kusdiana, Fuel 2001, 80, 225. [217] M. N. Varma, G. Madras, Ind. Eng. Chem. Res. 2007, 46, 1. [218] G. Madras, C. Kolluru, R. Kumar, Fuel 2004, 83, 2029. [219] A. Demirbas, Energy Sources Part A 2008, 30, 1645. [220] F. R. Abreu, M. B. Alves, C. C. S. Macedo, L. F. Zara, P. A. Z. Suarez, J. Mol. Catal. A: Chem. 2005, 227, 263. [221] C. S. Hsu, Practical Advances in Petroleum Processing (Ed.: P. R. Robinson), Springer, New York, 2006. [222] A. W. Schwab, G. J. Dykstra, E. Selke, S. C. Sorenson, E. H. Pryde, J. Am. Oil Chem. Soc. 1988, 65, 1781. [223] R. A. Niehaus, C. E. Goering, J. Savage, S. C. Sorenson, Trans. Am. Soc. Agr. Eng. 1986, 29, 683.

ChemSusChem 2009, 2, 278 300

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

299

S. Miertus et al.
[224] A. S. Ramadhas, S. Jayaraj, C. Muraleedharan, Renewable Energy 2004, 29, 727. [225] P. B. Weisz, W. O. Haag, P. G. Rodewald, Science 1979, 206, 57. [226] C. C. Chang, S. W. Wan, Ind. Eng. Chem. 1942, 42, 1543. [227] A. Crossley, T. D. Heys, B. J. F. Hudson, J. Am. Oil Chem. Soc. 1962, 39, 9. [228] J. W. Alencar, P. B. Alves, A. A. Craveiro, J. Agric. Food Chem. 1983, 31, 1268. [229] J. T. Kloprogge, L. V. Duong, R. L. Frost, Environ. Geol. 2005, 47, 967. [230] H. Tian, C. Li, C. Yang, H. Shan, Chin. J. Chem. Eng. 2008, 16, 394. [231] F. A. Twaiq, N. A. M. Zabidi, S. Bhatia, Ind. Eng. Chem. Res. 1999, 38, 3230. [232] D. Pioch, P. Lozano, R. C. Rasoanantoandro, J. Graille, P. Geneste, A. Guida, Oleagineux 1993, 48, 289. [233] I. C. P. Fortes, P. J. Baugh, J. Anal. Appl. Pyrolysis 1994, 29, 153. [234] D. G. B. Boocock, S. K. Konar, A. Mackay, P. T. C. Cheung, J. Liu, Fuel 1992, 71, 1291. [235] Y. S. Ooi, S. Bhatia, Microporous Mesoporous Mater. 2007, 102, 310. [236] Y. S. Ooi, R. Zakaria, A. R. Mohamed, S. Bhatia, Energy Fuels 2005, 19, 736. [237] Y. S. Ooi, R. Zakaria, A. R. Mohamed, S. Bhatia, Catal. Commun. 2004, 5, 441. [238] J. D. Adjaye, N. N. Bakhshi, Fuel Process. Technol. 1995, 45, 185. [239] S. P. R. Katikaneni, J. D. Adjaye, N. N. Bakhshi, Energy Fuels 1995, 9, 1065. [240] R. K. Sharma, N. N. Bakhshi, Energy Fuels 1993, 7, 306. [241] X. Dupain, D. J. Costa, C. J. Schaverien, M. Makkee, J. A. Moulijn, Appl. Catal. B 2007, 72, 44. [242] S. M. Sadrameli, A. E. S. Green, J. Anal. Appl. Pyrol., DOI: 10.1016/ j.jaap.2008.02.004. [243] G. W. Huber, P. OConnor, A. Corma, Appl. Catal. A: Gen. 2007, 329, 120. [244] I. Sebos, A. Matsoukas, V. Apostolopoulos, N. Papayannakos, Fuel 2009, 88, 145. [245] G. N. da Rocha Filho, D. Brodzki, G. Djega-Mariadassou, Fuel 1993, 72, 543. [246] A. A. Lappas, S. Bezergianni, I. A. Vasalos, Catal. Today, DOI: 10.1016/ j.cattod.2008.07.001. [247] G. W. Huber, A. Corma, Angew. Chem. 2007, 119, 7320; Angew. Chem. Int. Ed. 2007, 46, 7184. [248] P. Mki-Arvela, I. Kubickova, M. Snare, K. Eranen, D. Y. Murzin, Energy Fuels 2007, 21, 30. [249] M. Stumborg, A. Wong, E. Hogan, Bioresour. Technol. 1996, 56, 13. [250] http://www.ifp.com. [251] http://www.novabiosource.com.

Received: December 12, 2008

300

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 278 300

You might also like