You are on page 1of 15

Energy Conversion and Management 49 (2008) 27272741

Contents lists available at ScienceDirect

Energy Conversion and Management


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

A critical review of bio-diesel as a vehicular fuel


Mustafa Balat *, Havva Balat
Sila Science and Energy Unlimited Company, University Mahallesi, Trabzon, Turkey

a r t i c l e

i n f o

a b s t r a c t
The use of vegetable oils as alternative fuels has been around for one hundred years when the inventor of the diesel engine Rudolph Diesel rst tested peanut oil, in his compression-ignition engine. In 1970, scientists discovered that the viscosity of vegetable oils could be reduced by a simple chemical process and that it could perform as diesel fuel in modern engine. Considerable efforts have been made to develop vegetable oil derivatives that approximate the properties and performance of the hydrocarbon-based diesel fuels. Bio-diesel is an alternative to petroleum-based fuels derived from vegetable oils, animal fats, and used waste cooking oil including triglycerides. Bio-diesel production is a very modern and technological area for researchers due to the relevance that it is winning everyday because of the increase in the petroleum price and the environmental advantages. Transesterication is the most common method and leads to monoalkyl esters of vegetable oils and fats, now called bio-diesel when used for fuel purposes. 2008 Elsevier Ltd. All rights reserved.

Article history: Received 16 August 2007 Accepted 24 March 2008 Available online 2 May 2008 Keywords: Vegetable oil Bio-diesel Fuel properties Transesterication Market development Economy

1. Introduction Bio-diesel is an alternative to petroleum-based fuels derived from vegetable oils, animal fats, and used waste cooking oil including triglycerides. Since the petroleum crises in 1970s, the rapidly increasing prices and uncertainties concerning petroleum availability, a growing concern of the environment and the effect of greenhouse gases during the last decades, has revived more and more interests in the use of vegetable oils as a substitute of fossil fuel [1]. Vegetable oils are widely available from various sources, and the glycerides present in the oils can be considered as a viable alternative for diesel fuel [2]. Vegetable oils have good heating power and provide exhaust gas with almost no sulfur and aromatic polycyclic compounds. Due to the fact that vegetable oils are produced from plants, their burning leads to a complete recyclable carbon dioxide (CO2) [3]. Vegetable oils can be used as fuels for diesel engines, but their viscosities are much higher than usual diesel fuel and require modications of the engines [4]. 2. Bio-diesel production from vegetable oils 2.1. Literature review The use of vegetable oils as alternative fuels has been around for one hundred years when the inventor of the diesel engine Ru-

* Corresponding author. Tel.: +90 462 871 3025; fax: +90 462 871 3110. E-mail address: mustafabalat@yahoo.com (M. Balat). 0196-8904/$ - see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.enconman.2008.03.016

dolph Diesel rst tested peanut oil, in his compression-ignition engine [5]. In the 1930s and 1940s vegetable oils were used as diesel fuels from time to time, but usually only in emergency situations [6]. In 1940 rst trials with vegetable oil methyl and ethyl esters were carried out in France and, at the same time, scientists in Belgium were using palm oil ethyl ester as a fuel for buses. Not much was done until the late 1970s and early 1980s, when concerns about high petroleum prices motivated extensive experimentation with fats and oils as alternative fuels [7]. Bio-diesel (mono alkyl esters) started to be widely produced in the early 1990s and since then production has been increasing steadily. In the European Union (EU), bio-diesel began to be promoted in the 1980s as a means to prevent the decline of rural areas while responding to increasing levels of energy demand. However, it only began to be widely developed in the second half of the 1990s [8]. Bio-diesel production is a very modern and technological area for researchers due to the relevance that it is winning everyday because of the increase in the petroleum price and the environmental advantages [9]. The most common way of producing bio-diesel is the transesterication of vegetable oils and animal fats. Transesterication is not a new process. It was conducted as early as 1853 by two scientists E. Duffy and J. Patrick. Since that time several studies have been carried out using different oils such as cotton seed [10], soybean [11,12], waste cooking [13], rapeseed [1416], sunower seed [17], winter rape [18], frying [19,20], different alcohols such as methanol [21], ethanol [22], buthanol [23] as well as different catalysts, homogeneous ones such as sodium hydroxide [22,24], potassium hydroxide [22,25], sulfuric acid [26], and supercritical uids or enzymes such as lipases [2729].

2728

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

2.2. Vegetable oils as diesel fuels Vegetable oils, also known as triglycerides, have the chemical structure given in Fig. 1 comprise of 98% triglycerides and small amounts of mono- and diglycerides [32]. Triglycerides are esters of three fatty acids and one glycerol. These contain substantial amounts of oxygen in its structure [33]. The fatty acids vary in their carbon chain length and in the number of double bonds [34]. Different types of vegetable oils have different types of fatty acids. Table 1 shows the empirical formula and structure of various fatty acids present in vegetable oils. The disadvantages of vegetable oils as diesel fuel are [35]: higher viscosity, lower volatility and the reactivity of unsaturated hydrocarbon chains. Vegetable oils have their own advantages [1]: rst of all, they are available everywhere in the world. Secondly, they are renewable as the vegetables which produce oil seeds can be planted year after year. Thirdly, they are greener to the environment, as they seldom contain sulfur element in them. In 1970, scientists discovered that the viscosity of vegetable oils could be reduced by a simple chemical process and that it could perform as diesel fuel in modern engine. Fundamentally, high viscosity appears to be a property at the root of many problems associated with direct use of vegetable oils as engine fuel [36]. The vegetable oils, as alternative engine fuels, are all extremely viscous with viscosities ranging from 10 to 20 times greater than that of petroleum-derived diesel fuel [37]. Viscosity of No. 2 diesel fuel is 2.7 mm2/s at 38 C [34]. Viscosities of vegetable oils and their methyl esters are given in Fig. 2. The injection and atomization characteristics of the vegetable oils are signicantly different than those of petroleum-derived diesel fuels, mainly as the result of their high viscosities [38]. Modern diesel engines have fuel-injection systems that are sensitive to viscosity changes. A way to avoid these problems is to reduce the viscosity of vegetable oil in order to improve its performance [1]. Known problems, probable cause and

Fig. 2. Viscosities of vegetable oils and their methyl esters (mm2/s) at 38 C (source: Ref. [32]).

potential solutions for using straight vegetable oil in diesels are shown in Table 2. Several experimental studies showed that vegetable oils can be used as alternative fuel for diesel engines. Some of these vegetable oils are as follows: hazelnut [39], sunower [40], rapeseed [41,42], cottonseed [43,44], frying [45], jojoba [46], and Jatropha curcas [47]. The effect of temperature on viscosities of Jatropha oil and various blends in the range 2575 C has been investigated by Pramanik [47]. Results of this study are shown in Fig. 3. The results show that the viscosity of Jatropha oil is higher than diesel oil at any temperature. 2.3. Derivatives of vegetable oils as diesel fuels Considerable efforts have been made to develop vegetable oil derivatives that approximate the properties and performance of the hydrocarbon-based diesel fuels. The problems with substituting triglycerides for diesel fuels are mostly associated with their high viscosities, low volatilities and polyunsaturated character [31]. The viscosity of vegetable oils, when used as diesel fuel, can be reduced in at least four different ways: (1) dilution with hydrocarbons (blending), (2) emulsication, (3) pyrolysis (thermal cracking), and (4) transesterication (alcoholysis). Transesterication is the most common method and leads to monoalkyl esters of vegetable oils and fats, now called bio-diesel when used for fuel purposes. Pyrolysis, cracking, or other methods of decomposition of vegetable oils to yield fuels of varying nature is an approach that accounts for a signicant amount of the literature in historic times [46]. 2.3.1. Dilution Vegetable oils may be used with dilution modication technique as an alternative diesel fuel [47]. Dilution is an additional possible solution to the viscosity problem of vegetable oils as discussed above [48]. Viscosity of vegetable oil can be lowered by blending with pure ethanol. 25% of sunower oil and 75% of diesel were blending as diesel fuel. The viscosity was 4.88 cSt at 40 C, while the maximum specied ASTM value is 4.0 cSt at 38 C. This mixture was not suitable for long-term use in a direct injection engine [35,49]. A study was carried out by using the dilution technique on the same frying oil [50]. Results with this technology have been mixed and engine problems similar to those found with neat vegetable oils as fuels were observed here also. A model on vegetable oil atomization showed

Fig. 1. Structure of a typical triglyceride molecule (source: Ref. [30]).

Table 1 Chemical structure of common fatty acids Name of fatty acid Lauric Myristic Palmitic Stearic Arachidic Behenic Lignoceric Oleic Linoleic Linolenic Erucle Source: Ref. [30]. Chemical name of fatty acids Dodecanoic Tetradecanoic Hexadecanoic Octadecanoic Eicosanoic Docosanoic Tetracosanoic cis-9-Octadecenoic cis-9,cis-12-Octadecadienoic cis-9,cis-l2,cis-15Octadecatrienoic cis-13-Docosenoic Structure (xx:y) 12:0 14:0 16:0 18:0 20:0 22:0 24:0 18:1 18:2 18:3 22:1 Formula C12H24O2 C14H28O2 C16H32O2 C18H36O2 C20H40O2 C22H44O2 C24H48O2 C18H34O2 C18H32O2 C18H30O2 C32H42O2

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741 Table 2 Known problems, probable cause and potential solutions for using straight vegetable oil in diesels Problem Short term 1. Cold weather starting 2. Plugging and gumming of lters, lines and injectors 3. Engine knocking Long term 4. Coking of injectors on piston and head of engine 5. Carbon deposits on piston and head of engine 6. Excessive engine wear Probable cause High viscosity, low cetane, and low ash point of vegetable oils Natural gums (phosphatides) in vegetable oil. Other ash Potential solution

2729

Pre-heat fuel prior to injection. Chemically alter fuel to an ester. Partially rene the oil to remove gums. Filter to 4 lm

Very low cetane of some oils. Improper injection timing

Adjust injection timing. Use higher compression engines. Pre-heat fuel prior to injection. Chemically alter fuel to an ester Heat fuel prior to injection. Switch engine to diesel fuel when operation at part load. Chemically alter the vegetable oil to an ester Heat fuel prior to injection. Switch engine to diesel fuel when operation at part load. Chemically alter the vegetable oil to an ester Heat fuel prior to injection. Switch engine to diesel fuel when operation at part load. Chemically alter the vegetable oil to an ester. Increase motor oil changes. Motor oil additives to inhibit oxidation Heat fuel prior to injection. Switch engine to diesel fuel when operation at part load. Chemically alter the vegetable oil to an ester. Increase motor oil changes. Motor oil additives to inhibit oxidation

High viscosity of vegetable oil, incomplete combustion of fuel. Poor combustion at part load with vegetable oils High viscosity of vegetable oil, incomplete combustion of fuel. Poor combustion at part load with vegetable oils High viscosity of vegetable oil, incomplete combustion of fuel. Poor combustion at part load with vegetable oils. Possibly free fatty acids in vegetable oil. Dilution of engine lubricating oil due to blow-by of vegetable oil Collection of polyunsaturated vegetable oil blow-by in crankcase to the point where polymerization occurs

7. Failure of engine lubricating oil due to polymerization Source: Ref. [6].

soybean oil, 5% 190-proof ethanol and 20% 1-butanol was evaluated in the 200 h EMA screening test. The fuel passed the 200 h EMA test, but carbon and lacquer deposits on the injector tips, in-take valves and tops of the cylinder liners were major problems. The SNI fuel performed better than a 25% blend of sunower oil in diesel oil. The engine performances were the same for a microemulsion of 53% sunower oil and the 25% blend of sunower oil in diesel [6]. A microemulsion fuel containing soybean oil, methanol, 2-octanol, and a cetane enhancer was the cheapest vegetable oil-based alternative diesel fuel ever to pass the EMA test [48].

Fig. 3. Effect of temperature on viscosity of Jatropha oil and various blends (J: Jatropha oil; D: Diesel fuel; source: Ref. [45]).

that blends of No. 2 Diesel fuel with vegetable oil should contain from 0% to 34% vegetable oil if proper atomization was to be achieved [48]. 2.3.2. Microemulsion Various derivatives such as microemulsions or blends of various vegetable oils with conventional fuel have been proposed as alternative fuels for diesel engines [3]. Microemulsions are isotropic, clear, or translucent thermodynamically stable dispersions of oil, water, surfactant, and often a small amphiphilic molecule, called co-surfactant [31,5153]. The formation of microemulsions (cosolvency) is one of the four potential solutions for solving the problem of vegetable oil viscosity. Microemulsions are dened as transparent, thermodynamically stable colloidal dispersions in which the diameter of the dispersed-phase particles is less than onefourth the wavelength of visible light. Microemulsion-based fuels are sometimes also termed hybrid fuels, although blends of conventional diesel fuel with vegetable oils have also been called hybrid fuels [48]. Some of these fuels in diesel engines were tested by the Engine Manufacturers Association (EMA). Shipp non-ionic (SNI) fuel containing 50% No. 2 Diesel fuel, 25% degummed and alkali-rened

2.3.3. Pyrolysis (thermal cracking) The pyrolysis refers to chemical change caused by the application of thermal energy in the presence of an air or nitrogen sparge [51]. The pyrolysis of fats has been investigated for more than 100 years, especially in those areas of the world that lack deposits of petroleum [6]. The pyrolysis of different triglycerides was used for fuel supply in different countries during the First and Second World Wars. For instance, a tung oil pyrolysis batch system was used in China as a hydrocarbon supply during World War II. These hydrocarbons were used as raw materials for gasoline and diesellike fuel production in a cracking system similar to the petroleum process now used [54]. Thermal decomposition of triglycerides produces the compounds of classes including alkanes, alkenes, alkadienes, aromatics and carboxylic acids. Different types of vegetable oils produce large differences in the composition of the thermally decomposed oil. Fig. 4 outlines a schematic that accounts for the formation of alkanes, alkenes, alkadienes, aromatics and carboxylic acids from pyrolysis of triglycerides [31]. The main components are alkanes and alkanes, which accounted for approximately 60% of the total feeder weight. Carboxylic acids accounted for another 9.616.1%. It is believed that as the reaction progresses the residue becomes less reactive and forms stable chemical structures, and consequently the activation energy increases as the decomposition level of vegetable oil residue increases [55]. The yields of pyrolysis of vegetable oils are given in Table 3. Soybean oil pyrolyzed distillate, which consisted mainly of alkanes, alkenes, and carboxylic acids had a cetane number of 43, exceeding that of soybean oil (37.9) and the ASTM minimum value

2730

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

Fig. 4. The mechanism of thermal decomposition of triglycerides (source: Ref. [31]).

Table 3 Yields of pyrolysis of vegetable oils (percent by weight) High oleic acid safower N2 sparge Alkanes Alkanes Alkadienes Carboxylic acids Unresolved unsaturates Aromatics Unidentied Source: Refs. [6,33,49]. 37.5 22.2 8.1 11.5 9.7 2.3 8.7 Air 40.9 22.0 13.0 176.1 10.1 2.2 12.7 Soybean oil N2 sparge 31.1 28.3 9.4 12.2 5.5 2.3 10.9 Air 29.9 24.9 10.9 9.6 5.1 1.9 12.6

2.3.4. Transesterication (alcoholysis) Transesterication is the reaction of vegetable oil or animal fat with an alcohol, in most cases methanol, to form esters and glycerol. The transesterication reaction is affected by alcohol type, molar ratio of glycerides to alcohol, type and amount of catalyst, reaction temperature, reaction time and free fatty acids and water content of vegetable oils or animal fats. The transesterication reaction proceeds with or without a catalyst by using primary or secondary monohydric aliphatic alcohols having 18 carbon atoms as follows [21,35,48]: Triglycerides Monohydric alcohol ! Glycerin Mono-alkyl esters 2

of 40. The viscosity of the distillate was 10.2 cSt at 38 C, which is higher than the ASTM specication for Diesel fuel No. 2 (1.9 4.1 cSt) but considerably below that of soybean oil (32.6 cSt). Short-term engine tests were carried out on this fuel. Used cottonseed oil from the frying process was decomposed with Na2CO3 as catalyst at 450 C to give a pyrolyzate containing mainly C820 alkanes (70%) besides alkenes and aromatics [48]. Many kinds of vegetable oil species have been subjected to pyrolysis conditions. Some of these vegetable oils are as follows: soybean [54,56,57], rapeseed [58], palm tree [54,59], castor [54], safower [58], olive husk [60], and tung [61]. Recently, the yields of decarboxylation products by pyrolysis from vegetable oil soaps have been investigated by Demirbas and Kara [32]. The maximum decarboxylation products of pyrolyses were 96.8%, 97.1%, 97.5%, and 97.8%, respectively, from sunower oil, corn oil, cottonseed oil, and soybean oil at 610 K (337 C), respectively. Oxidative pyrolysis of Na-soaps is given as following reaction [32,33,49]: 4RCOONa O2 ! 2R-R 2Na2 CO3 2CO2 1

Generally, the reaction temperature near the boiling point of the alcohol is recommended. Nevertheless, the reaction may be carried out at room temperature [62]. The reactions take place at low temperatures ($65 C) and at modest pressures (2 atm, 1 atm = 101.325 kPa). Bio-diesel is further puried by washing and evaporation to remove any remaining methanol. The oil (87%), alcohol (9%), and catalyst (1%) are the inputs in the production of bio-diesel (86%), the main output [63]. Pre-treatment is not required if the reaction is carried out under high pressure (9000 kPa) and high temperature ($240 C), where simultaneous esterication and transesterication take place with maximum yield obtained at temperatures ranging from 60 to 80 C at a molar ratio of 6:1 [30]. The alcohols employed in the transesterication are generally short chain alcohols such as methanol, ethanol, propanol, and butanol. It was reported that when transesterication of soybean oil using methanol, ethanol and butanol was performed, 9698% of ester could be obtained after 1 h of reaction [2]. 2.3.4.1. Catalytic transesterication. Catalysts used for the transesterication of triglycerides are classied as alkali, acid, enzyme or heterogeneous catalysts, among which alkali catalysts like sodium hydroxide, sodium methoxide, potassium hydroxide, potassium methoxide are more effective [5]. Sulfuric acid, hydrochloric acid and sulfonic acid are usually preferred as acid catalysts. The heterogeneous catalysts include enzymes, titanium-silicates,

The soaps obtained from the vegetable oils can be pyrolyzed into hydrocarbon-rich products according to Eq. (1) with higher yields at lower temperatures.

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

2731

alkaline-earth metal compounds, anion exchange resins and guanadines heterogenized on organic polymers [64]. The catalytic transesterication of vegetable oils with methanol is an important industrial method used in bio-diesel synthesis. Also known as methanolysis, this reaction is well studied and established using acids or alkalis, such as sulfuric acid or sodium hydroxide as catalysts. However, these catalytic systems are less active or completely inactive for long chain alcohols. Usually, industries use sodium or potassium hydroxide or sodium or potassium methoxide as catalyst, since they are relatively cheap and quite active for this reaction [65]. Fig. 5 shows the catalytic bio-diesel production diagram. 2.3.4.1.1. Alkali-catalyzed transesterication. The transesterication process is catalyzed by alkaline metal alkoxides [53,66], and hydroxides [2,3,6770], as well as sodium or potassium carbonates [71,72]. A simplied block ow diagram for a typical base-catalyzed process for the production of bio-diesel is shown in Fig. 6 [73]. The alkaline catalysts show high performance for obtaining vegetable oils with high quality, but a question often arises; that is, the oils contain signicant amounts of free fatty acids which cannot be converted into bio-diesels but to a lot of soap [74]. These free fatty acids react with the alkaline catalyst to produce soaps that inhibit the separation of the bio-diesel, glycerin, and wash water [75]. Triglycerides are readily transesteried batchwise in the presence of alkaline catalyst at atmospheric pressure and at a temperature of approximately 6070 C with an excess of methanol [31]. It often takes at least several hours to ensure the alkali

Fig. 5. Catalytic bio-diesel production diagram.

(NaOH or KOH) catalytic transesterication reaction is complete. Moreover, removal of these catalysts is technically difcult and brings extra cost to the nal product [49,76]. Alkaline metal alkoxides (as CH3ONa for the methanolysis) are the most active catalysts, since they give very high yields (>98%) in short reaction times (30 min) even if they are applied at low molar concentrations (0.5 mol%). Alkaline metal hydroxides (KOH and NaOH) are cheaper than metal alkoxides, but less active. Nevertheless, they are a good alternative since they can give the same high conversions of vegetable oils just by increasing the catalyst concentration to 1 or 2 mol% [77]. The great thing about this transesterication process is that some of the methanol can be recovered and that glycerin (that is used in pharmaceuticals and other applications) is also a by-product. In this process, the glycerin needs to be removed to assure that it does not get converted to formaldehyde or acetaldehyde when burned; both of which would pose a health hazard. It is particularly interesting that the oil to be used is typically derived from the soybean, a high protein legume of which 80% is animal feed and the remaining 20% will serve as the oil in the bio-diesel synthesis [78]. 2.3.4.1.2. Acid catalyzed transesterication. The transesterication process is catalyzed by sulfuric [7981], hydrochloric [79,82], and organic sulfonic acids [83]. A simplied block ow diagram of the acid-catalyzed process is shown in Fig. 7. In general, acid catalyzed reactions are performed at high alcohol-to-oil molar ratios, low-tomoderate temperatures and pressures, and high acid catalyst concentrations. Acid-catalyzed reactions require the use of high alcohol-to-oil molar ratios in order to obtain good product yields in practical reaction times. However, ester yields do not proportionally increase with molar ratio. For instance, for soybean methanolysis using sulfuric acid, ester formation sharply improved from 77% using a methanol-to-oil ratio of 3.3:1 to 87.8% with a ratio of 6:1. Higher molar ratios showed only moderate improvement until reaching a maximum value at a 30:1 ratio (98.4%) [73]. Despite its insensitivity to free fatty acids in the feedstock, acid-catalyzed transesterication has been largely ignored mainly because of its relatively slower reaction rate [84]. 2.3.4.1.3. Enzyme catalyzed transesterication. The transesterication process is catalyzed by lipases such as Candida antarctica

Fig. 6. A simplied block ow diagram for a typical base-catalyzed process for the production of bio-diesel (source: Ref. [73]).

2732

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

Fig. 7. A simplied block ow diagram of the acid-catalyzed process for the production of bio-diesel (source: Ref. [73]).

[10,11,28,85], Candida rugasa [86], Pseudomonas cepacia [87,88], immobilized lipase (Lipozyme RMIM) [89,90], Pseudomonas sp. [91], and Rhizomucor miehei [91,92]. The enzymatic alcoholysis of soybean oil with methanol and ethanol was investigated using a commercial, immobilized lipase (Lipozyme RMIM) [90]. In that study, the best conditions were obtained in a solvent-free system with ethanol/oil molar ratio of 3.0, temperature of 50 C, and enzyme concentration of 7.0% (w/ w). They obtained yield 60% after 1 h of reaction. In another study, Shah and Gupta [88] obtained a high yield (98%) by using P. cepacia lipase immobilized on celite at 50 C in the presence of 45% (w/w) water in 8 h. As for the enzyme-catalyzed system, it requires a much longer reaction time than the other two systems [84]. Both extracellular and intracellular lipases are also able to effectively catalyze the transesterication of triglycerides in either aqueous or non-aqueous systems, and as shown in Table 4, enzymatic transesterication methods can overcome the problems mentioned above [51]. The advantages of lipase-catalyzed transesterication, compared to the chemically-catalyzed reaction, are emphasized [93]. The main problem of the lipase-catalyzed process is the high cost of the lipases used as catalyst [10]. The enzyme reactions are highly specic and chemically clean. Because the alcohol can be inhibitory to the enzyme, a typical strategy is to feed the alcohol

into the reactor in three steps of 1:1 mole ratio each. The reactions are very slow, with a three step sequence requiring from 4 to 40 h, or more. The reaction conditions are modest, from 35 to 45 C [94]. 2.3.4.2. Non-catalytic supercritical methanol transesterication. The transestercation of triglycerides by supercritical methanol (SCM), ethanol, propanol and butanol has proved to be the most promising process [35]. Fig. 8 shows supercritical ethanol transesterication system. The critical temperatures and critical pressures of the various alcohols are shown in Table 5. Recently, Saka and Kusdiana [95] have developed a catalyst-free method for biodiesel fuel production by employing supercritical methanol. The supercritical treatment at 350 C, 43 MPa and 240 s with a molar

Table 4 Comparison between alkali-catalysis and lipase-catalysis methods for bio-diesel fuel production Alkali-catalysis process Reaction temperature (C) Free fatty acids in raw materials Water in raw materials Yield of methyl esters Recovery of glycerol Purication of methyl esters Production of catalyst Source: Ref. [51]. 6070 Saponied products Interference with the reaction Normal Difcult Repeated washing Cheap Lipase-catalysis process 3040 Methyl esters No inuence Higher Easy None Relatively expensive

Fig. 8. Supercritical ethanol transesterication system. 1. Autoclave, 2. Electrical furnace, 3. Temperature control monitor, 4. Pressure control monitor, 5. Product exit valve, 6. Condenser, 7. Product collecting vessel (source: Ref. [97]).

Table 5 Critical temperatures and critical pressures of various alcohols Alcohol Methanol Ethanol 1-Propanol 1-Butanol Source: Ref. [35]. Critical temperature (C) 239.2 243.2 264.2 287.2 Critical pressure (MPa) 8.1 6.4 5.1 4.9

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

2733

ratio of 42 in methanol is the optimum condition for transesterication of rapeseed oil to bio-diesel fuel [96]. Results of this study are given Table 6. Demirbas [76] investigated the changes in yield percentage of methyl esters with supercritical methanol method with a molar ratio of 41. The critical temperature and the critical pressure of methanol were 512.4 K and 8.0 MPa, respectively. In that study, it was concluded that increasing reaction temperature, especially supercritical temperatures had a favorable inuence on ester conversion. The yields of ethyl esters from vegetable oils via transesterication in supercritical ethanol were investigated [97]. Fig. 9 shows the changes in yield percentage of ethyl esters as treated with sub- and supercritical ethanol at different temperatures as a function of reaction time. The critical temperature and the critical pressure of ethanol were 516.2 K and 6.4 MPa, respectively. The non-catalyst options are designed to overcome the reaction initiation lag time caused by the extremely low solubility of the alcohol in the TG phase. One approach that is nearing commercialization is the use of a co-solvent, tetrahydrofuran (THF), to solubilize the methanol. The result is a fast reaction, on the order of 5 10 min, and no catalyst residues in either the ester or the glycerol phase. The THF co-solvent is chosen, in part, because it has a boiling point very close to that of methanol. This system requires a rather low operating temperature, 30 C. Another non-catalytic approach is the use of a high (42:1) alcohol to oil ratio. Under supercritical conditions (350400 C and >80 atm or 1200 psi). The reaction is complete in about 4 min [94]. Reaction by supercritical methanol has some advantages [98]: (1) glycerides and free fatty acids are reacted with equivalent rates, (2) the homogeneous phase eliminates diffusive problems, (3) the process tolerates great percentages of water in the feedstock catalytic process require the periodical renoval of water in the feedstock or in intermediate stage to prevent catalyst deactivation, (4) the catalyst renoval step is eliminated, and (5) if high methanol:oil ratios are used, total conversion of the oil can be achieved in a few minutes. Some disadvantages of the one-stage supercritical method are clear: (1) it operates at very high pressures (25 40 MPa), (2) the high temperatures bring along proportionally high heating and cooling costs, (3) high methanol:oil ratios (usually set at 42) involve high costs for the evaporation of the unreacted methanol, and (4) the process as posed to date does not explain how to reduce free glycerol to less than 0.02% as established in the ASTM D6584 or other equivalent international standards. 2.3.4.3. Effect of reaction parameters on conversion yield of transesterication. 2.3.4.3.1. Effect of free fatty acid and moisture. The free fatty acid (FFA) and moisture content are key parameters for determining the viability of the vegetable oil transesterication process [5]. In the transesterication, FFAs and water always produce negative effects, since the presence of FFAs and water causes soap formation, consumes catalyst and reduces catalyst effectiveness, all of which result in a low conversion [99]. In catalyzed methods, the presence of water has negative effects on the yields of methyl esters [35]. In

Fig. 9. Changes in yield percentage of ethyl esters as treated with sub- and supercritical ethanol at different temperatures as a function of reaction time. Molar ratio of vegetable oil-to-ethyl alcohol: 1:40 (source: Ref. [97]).

acid catalyzed transesterication, fatty acids can be formed by the reaction of carbocation II with the water in the reaction mixture [100]. These free fatty acids react with the alkaline catalyst to produce soaps that inhibit the separation of the bio-diesel, glycerin, and wash water [75]. To carry the base catalyzed reaction to completion; a free fatty acid value lower than 3% is needed [5]. Figs. 10 and 11 show a direct comparison of the yield of methyl esters from various preparation methods as triglycerides with various water and FFA contents are treated [35,101,102]. The presence of water has a greater negative effect on transesterication than that of the FFAs. Ma et al. [103] investigated the transesterication of beef tallow catalyzed by sodium hydroxide (NaOH) in presence of FFAs and water. These authors reported that water and free fatty acid contents of beef tallow had to be maintained below 0.06 wt% and 0.5 wt%, respectively. Using NaOH catalyzed transesterication, methyl esters can generally be prepared in high yields for low FFA oils, being nearly quantitative for the palm oils containing <1% FFA. For example, the yield of methyl esters from RBD (rened, bleached and deodorized) palm oil with about 0.05% FFA was 98% [100]. 2.3.4.3.2. The effect of molar ratio and alcohol type. One of the most important factors that affect the yield of ester is the molar ratio of alcohol to triglyceride. Although the stoichiometric molar ratio of methanol to triglyceride for transesterication is 3:1, higher molar ratios are used to enhance the solubility and to increase the contact between the triglyceride and alcohol molecules [104]. Higher molar ratios result in greater ester conversion in a shorter time. In the

Table 6 Comparison of the yields in alkaline-catalyzed, acid-catalyzed and supercritical methanol Raw material Free fatty acid content (wt%) Water content (wt%) Yields of methyl esters (wt%) Alkaline-catalyzed Rapeseed oil Palm oil Frying oil Waste palm oil Source: Ref. [96]. 2.0 5.3 5.6 >20.0 0.02 2.1 0.2 >61.0 97.0 94.4 94.1 Acid-catalyzed 98.4 97.8 97.8 Supercritical methanol 98.5 98.9 96.9 95.8

2734

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

Fig. 10. Yields of methyl esters as a function of water content in transesterication of triglycerides (source: Refs. [35,101,102]).

Fig. 11. Yields of methyl esters as a function of free fatty acid content in transesterication of triglycerides (source: Refs. [35,101,102]).

transesterication of peanut oil with ethanol, a 6:1 molar ratio liberated signicantly more glycerine than a ratio of 3:1 [105]. Freedman et al. [106] investigated the effect of molar ratio on the transesterication of sunower oil with methanol. They obtained the results for transesterication of sunower oil, in which the molar ratio varied from 6:1 to 1:1 and concluded that 98% conversion to ester was obtained at a molar ratio of 6.1. The transesterication of Rice Bran oil with methanol was studied at molar ratios of 4:1, 5:1, and 6:1 [107]. At molar ratios of 4:1 and 5:1, there was signicant increase in yield when the reaction time was increased from 4 to 6 h. Among the three molar ratios studied, ratio 6:1 gave the best results. Results obtained are summarized in

Table 7. In the previously mentioned work of Noureddini et al. [104], the authors showed that at all mixing levels the molar ratio of 8:l results in signicantly higher conversions than 6:1. Tomasevic and Siler-Marinkovic [20] observed that the molar ratio is much more effective than catalyst on the transesterication reaction. The molar ratio is associated with the type of catalyst used. Acid-catalyzed reactions require the use of high alcohol-to-oil molar ratios in order to obtain good product yields in practical reaction times. However, ester yields do not proportionally increase with molar ratio. For instance, for soybean methanolysis using sulfuric acid, ester formation sharply improved from 77% using a methanol-to-oil ratio of 3.3:1 to 87.8% with a ratio of 6:1. Higher molar ratios showed only moderate improvement until reaching a maximum value at a 30:1 ratio (98.4%) [73]. The effect of molar ratio of dibutyltin oxide to dimethyl carbonate (DMC) on the transesterication of DMC with phenol was studied in the presence of triic acid (Bu2SnO/CF3SO3H = 1/1) at 180 C. The yields of methyl phenyl carbonate (MPC) and diphenyl carbonate (DPC) increased rapidly with increasing molar ratio of dibutyltin oxide/DMC = 0.02. Above 0.02, however, the yields of MPC and DPC remained unchanged. Results of this study are shown in Fig. 12 [108]. Another important variable affecting the yield of methyl ester is the type of alcohol to triglyceride. In general, short chain alcohols such as methanol, ethanol, propanol, and butanol can be used in the transesterication reaction to obtain high methyl ester yields. Canakci and Van Gerpen [109] investigated the effect of different alcohol types on acid-catalyzed transesterication of pure soybean oil. They obtained yields from 87.8% to 95.8% after 48 and 96 h of reaction. Results obtained are summarized in Table 8. 2.3.4.3.3. The effect of catalyst. Catalysts used for the transesterication of triglycerides are classied as alkali, acid, enzyme. Alkalicatalyzed transesterication is much faster than acid-catalyzed transesterication and is most often used commercially [6]. May [100] studied the effect of different catalysts types on methanolysis of RBD palm oil with a low FFA content of <0.1%. In that study, it was concluded that Na, NaOH and KOH are effective catalysts. Results of this study are shown in Table 9. In the previously mentioned work of Stavarache et al. [3], the authors investigated the effect of different catalyst concentrations on base-catalyzed transesterication during bio-diesel production from vegetable oil by means of ultrasonic energy. The best yields were obtained when the catalyst was used in small concentration, i.e., 0.5% wt/wt of oil. Meneghetti et al. [110] investigated the effect of different catalyst types at different temperatures during the production of free and bound ethyl ester (FAEE) from castor oil. Results from that study showed that hydrochloric acid is much more effective than sodium hydroxide at higher reaction temperatures (Fig. 13). 2.3.4.3.4. The effect of reaction temperature and time. Transesterication can occur in different temperatures depending on the type

Table 7 Effect of reaction time and methanol-to-oil molar ratio Serial no. Reaction time (min) Percent conversion at molar ratio 4:1 1 2 3 4 5 6 7 Source: Ref. [107]. 15 30 45 60 120 240 360 85 86 88 90 91 92 92 5:1 89 89 91 92 93 94 95 6:1 93 94 94 94 94 95 97 Percent yield at molar ratio 4:1 45 53 57 61 66 66 83 5:1 46 54 59 67 67 67 82 6:1 62 66 72 77 72 82 83

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

2735

Fig. 12. Effect of molar ratio on the transesterication of DMC with phenol. Dimethyl carbonate (40 mmol), phenol (200 mmol), benzene (40 ml), molecular sieves (30 g), Bu2SnO/CF3SO3H = 1/1, 180 C, 3 h (source: Ref. [108]).

Fig. 13. The effect of different catalyst types at different temperatures during the production of FAEE from castor oil (percentage yield) (source: Ref. [110]).

Table 8 The effect of alcohol type on conversion and specic gravity of ester Alcohol type Boiling point (C) 65 82.4 117 78.5 Reaction temperature (C) 60 75 110 75 Ester conversion (C) 87.8 92.9 92.1 95.8 Specic gravity of bio-diesel 0.8876 0.8786 0.8782 0.8814

Methanol 2-Propanol 1-Butanol Ethanol Source: Ref. [109].

tion temperature on methyl ester yield of non-catalytic transesterication in sub- and supercritical methanol from sunower oil. In that study, it was concluded that increasing the reaction temperature had a favorable inuence on the yield of methyl ester (Fig. 14). The conversion rate increases with reaction time. Kim et al. [24] studied the transesterication reaction, with methanol (6:1), of vegetable oils using NaOH and Na/NaOH/c-Al2O3 as catalysts at 60 C. They reached the maximum bio-diesel yield within 1 h both for the case of homogeneous and heterogeneous catalyst system. For the homogeneous catalyst system, the maximum bio-diesel production yield was higher by 20% than that of the heterogeneous catalyst system [24]. Results of this study are shown in Fig. 15. 3. Source for bio-diesel production Bio-diesel can be produced from any material that contains fatty acids, be they linked to other molecules or present as free fatty acids. Thus various vegetable fats and oils, animal fats, waste greases, and edible oil processing wastes can be used as feedstocks for bio-diesel production [117]. There are more than 350 oil-bearing crops identied, among which only sunower, safower, soybean, cottonseed, rapeseed, and peanut oils are considered as potential alternative fuels for diesel engines [32,35,97,118,119]. However, in recent years, animal fats and especially recycled greases and used vegetable oils have found increasing attention

Table 9 Effect of different catalysts on transesterication of palm oila Catalyst Na NaOH KOH H2SO4 (conc.) HCl (conc.) Ion exchange resin (H+) Dowex 50 (Na+) Acid treated orisil Activated silica gel Amount (wt% based on oil) 0.1 0.2 1 1 1 2 1 2 1 Reaction time (min) 1632 1632 1632 >300 >300 >300 >300 >300 >300 Remarks 99% yield 98% yield 98% yield 50% yield 30% yield Too slow Too slow Too slow Not suitable

a The following conditions were used: type of oil = RBD palm oil (FFA = 0.05%); ratio of oil-to-solvent (methanol) = 1:15.6; temperature = reux temperature. (Source: Ref. [100].)

of oil employed [5,6,111]. A few works reported the reaction at room temperature [22,112114]. Encinar et al. [22] studied the transesterication reaction, with ethanol, of Cynara cardunculus L. oils using sodium and potassium hydroxides as catalysts and reported 91.6% conversion at room temperature. For the transesterication of palm oil with methanol (6:1) and 1% KOH, the reaction was studied at different temperatures [115]. After 4 min, ester yields were 73% and 82% for 50 and 65 C, respectively. The effect of reaction temperature on production of propyl oleate was examined at the temperature range from 40 C to 70 C with free immobilized P. uorescens lipase [116]. The conversion ratio to propyl oleate was observed highest at 60 C, whereas the activity highly decreased at 70 C. Demirbas [21] investigated the effect of reac-

Fig. 14. Effect of temperature on methyl ester yield of non-catalytic transesterication in sub- and supercritical methanol from sunower oil (source: Ref. [21]).

2736

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

(Fig. 16) [111]. In the EU bio-diesel is predominantly produced from vegetable oils and especially from rapeseed and sunower. The majority of bio-diesel in the United States is made from soybean oil (estimated at over 90%), but a few producers use other oilseed crops, palm oil, animal fats and recycled oils to make biodiesel [122]. Different types of oils have different types of fatty acids (Table 1). The fatty acids are different in relation to the chain length, degree of unsaturation or presence of other chemical functions (Table 10) [111]. 4. Bio-diesel market development The global bio-diesel industry is among the fastest-growing markets the chemical industry has ever seen [123]. Bio-diesel production is highest in the EU, where more bio-diesel is produced than ethanol, but total production of both fuels is fairly small compared to production of bio-ethanol in Brazil and the United States [124]. Two factors have contributed to the EUs becoming the world leader in bio-diesel production. One is the reform of the Common Agricultural Policy (CAP), a supranational and domestically oriented farm policy for EU member countries, adopted in 1992 and implemented in 19931994. The 1992 reform addressed agricultural surpluses through supply control in the form of a mandatory, paid, set-aside program. The second factor is high fuel taxes, which have enabled indirect subsidies for bio-fuel production through partial or full exemption of the fuel excise tax. Taxes normally constitute 50% or more of the retail price of diesel in EU member states. In February 1994, the European Parliament adopted a 90% tax exemption for bio-diesel [121]. The EU is the world leader in the production and consumption of bio-diesel. According to the European Bio-diesel Boards 2007 gures [125], the EU produced 4.98 million tons (approx. 1.5 billion gallons) of bio-diesel in 2006, up from 3.18 million tons (approx. 961 million gallons) in 2005. The EU accounted for 77% of global bio-diesel production in 2006. Bio-diesel accounted for nearly 80% of EU bio-fuel production [126]. Table 11 shows the EUs bio-diesel production during 20042006 [125,127]. The EU is expected to continue to be the main market and producer of bio-diesel, followed by the United States and Brazil [8]. The EC Bio-fuels Directive of 2003 required a voluntary market share of 5.75% bio-fuels for each member state by 2010. Other countries (including Netherlands, India, China, Thailand and New Zealand) and individual states in the United States and Canada have since established mandatory bio-fuels targets and yet others have removed excise taxes. Such policies should see additional capacity for bio-ethanol and bio-diesel production developed [128]. The potential market for bio-diesel is estimated to be in the order of 20 EJ

Fig. 15. Effect of reaction time on the bio-diesel production yield. Methanol/VO molar ratio 6:1, reaction temperature 60 C, stirring speed 300 rpm, without cosolvent (source: Ref. [24]).

as sources of bio-diesel, the latter primarily as inexpensive feedstocks [120]. Rapeseed and soybeans are the two most commonly used feedstocks for bio-diesel today. Rapeseed has a higher oil content than soybeans. They both have alternative markets as cooking oil and food, pushing up their prices. Rapeseed produces about 435 l of bio-diesel per acre, but high-yield rapeseed can produce as much as 550 l. Soybeans produce about 160 l per acre. Sunowers can produce 280 l of bio-diesel per acre, palm oil as much as 975 l [121]. Any fatty acid source may be used to prepare bio-diesel, but most scientic articles take soybean as a bio-diesel source

Fig. 16. Leading bio-diesel sources cited in scientic articles (source: Ref. [111]).

Table 10 Fatty acid composition of some vegetable oils (%) Vegetable oil Tallow Coconut oil Olive oil Groundnut oil Cotton oil Corn oil Soybean oil Hazelnut kernel Poppy seed Rapeseed Safower seed Sunower seed Castor oil Source: Ref. [111]. Palmitic 16:0 29.0 5.0 14.6 8.5 28.6 6.0 11.0 4.9 12.6 3.5 7.3 6.4 Stearic 18:0 24.5 3.0 6.0 0.9 2.0 2.0 2.6 4.0 0.9 1.9 2.9 3.0 Palmitoleic 16:1 0.1 0.2 0.1 0.1 0.1 0.1 3.0 Oleic 18:1 44.5 6.0 75.4 51.6 13.0 44.0 20.0 81.4 22.3 54.1 13.5 17.7 3.0 Linoleic 18:2 10.0 26.0 57.2 48.0 64.0 10.5 60.2 22.3 77.0 72.8 1.2 Ricinic 12-OH-oleic 89.5 Other acids 65.0 0.2 3.0 0.3 0.8 9.1 0.2 0.1 0.3

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741 Table 11 The EUs bio-diesel production by country (1000 tons) Country Germany France Italy UK Austria Poland Czech Rep. Spain Portugal Slovakia Denmark Others Total EU Source: Refs. [125,127]. 2004 1035 348 320 9 57 0 60 13 0 15 70 6 1933 2005 1669 492 396 51 85 100 133 73 1 78 71 35 3184 2006 2662 743 447 192 123 116 107 99 91 82 80 148 4890

2737

by 2050, assuming development of synthetic bio-fuel production technologies [129]. Bio-diesel output in Germany where the biggest in the EU industry was hit by taxes last year jumped to 2.66 million tons in 2006 from 1.67 million tons in 2005 (Table 11). Looking at world bio-diesel production, 42% of the world total in 2006 was accounted for by Germany, where market conditions were highly favorable (Fig. 17). Other countries with signicant bio-diesel markets in 2006 included France, the United States, Italy and the United Kingdom. The United States was the worlds second-largest producer of bio-diesel, behind the EU. The United States bio-diesel production remained very small and at until the United States Department of Agriculture (USDA) created the Bioenergy Program in Fiscal Year (FY) 2000 that encouraged bio-diesel production through cash payments to producers [122]. Mostly as a result of this program, bio-diesel production reached from 1 million gallons (approx. 3300 tons) in 1999 [126] to 250 million gallons (approx. 836,000 tons) in 2006 [125]. By 2010, the United States is expected to become the largest single bio-diesel market, accounting for roughly 18% of world bio-diesel consumption [123]. Experts predict that, in the best case scenario, in the next 20 years bio-diesel could take care of 25% of the United States diesel needs [8]. 5. Bio-diesel as a vehicular fuel Bio-diesel is a cleaner-burning diesel replacement fuel made from natural, renewable sources such as new and used vegetable oils and animal fats. Just like petroleum diesel, bio-diesel operates in compression-ignition engines or Diesel engines. Bio-diesel has physical properties very similar to conventional diesel [76]. The bio-diesel was characterized by determining its density, viscosity, high heating value, cetane index, cloud and pour points, character-

istics of distillation, and ash and combustion points according to ISO norms [22,130]. Selected properties of diesel and bio-diesel fuels are given in Table 12 [131]. Viscosity is the most important property of bio-diesel since it affects the operation of the fuel injection equipment, particularly at low temperatures when the increase in viscosity affects the uidity of the fuel. Bio-diesel has a viscosity close to that of diesel fuels [21,35]. The higher viscosity range of bio-diesel helps to reduce barrel/plunger leakage and increase injector efciency in engines [121]. Viscosity measurements have been made over the temperature range 20100 C for blends of different bio-diesel with No. 2 diesel fuel [131]. The viscosity of the distillate was 10.2 mm2/s at 38 C, which is higher than the ASTM specication for No. 2 Diesel fuel (1.94.1 mm2/s) but considerably below that of soybean oil (32.6 mm2/s) [101]. Bio-diesel offers safety benets over petroleum diesel because it is much less combustible, with a ash point greater than 150 C, compared to 77 C for petroleum diesel [132]. The cetane number of bio-diesel is generally higher than conventional diesel. The cetane number of bio-diesel is around 50. Bio-diesel has lower volumetric heating values (about 12%) than diesel fuels but has a high cetane number and ash point [31]. Density is another important property of bio-diesel. It is the weight of a unit volume of uid. Specic gravity is the ratio of the density of a liquid to the density of water. Specic gravity of bio-diesels ranges between 0.87 and 0.89. Fuel injection equipment operates on a volume metering system, hence a higher density for bio-diesel results in the delivery of a slightly greater mass of fuel [35,133]. Two important parameters for low temperature applications of a fuel are Cloud Point (CP) and Pour Point (PP). The CP is the temperature at which wax rst becomes visible when the fuel is cooled. The PP is the temperature at which the amount of wax out of solution is sufcient to gel the fuel, thus it is the lowest temperature at which the fuel can ow. Bio-diesel has higher CP and PP compared to conventional diesel [35,134]. The esters have CP and PP that are 1525 C higher than those of diesel fuels [31]. Methyl esters of vegetable oils have several outstanding advantages among other new-renewable and clean engine fuel alternatives [21,102,135]. A number of technical advantages of biodiesel fuel [130,136]: (1) it prolongs engine life and reduces the need for maintenance (bio-diesel has better lubricating qualities than fossil diesel); (2) it is safer to handle, being less toxic, more biodegradable, and having a higher ash point; (3) it reduces some exhaust emissions (although it may, in some circumstances, raise

Table 12 Specications of diesel and bio-diesel fuels Fuel property Fuel standard Fuel composition Lower heating value (MJ/m3) Kinematic viscosity at 40 C (mm2/s) Specic gravity at 15.5 C Density at 15 C (kg/m3) Water (ppm by wt.) Carbon (wt%) Hydrogen (wt%) Oxygen (by diff.) (wt%) Sulfur (wt%) Boiling point (C) Flash point (C) Cloud point (C) Pour point (C) Cetane number Stoichiometric air/fuel ratio (wt./wt.) Diesel ASTM D975 C10C21 HC 36.6 103 1.34.1 0.85 848 161 87 13 0 0.05 max 188343 6080 15 to 5 35 to 15 4055 15 Bio-diesel ASTM PS 121 C12C22 FAME 32.6 103 1.96.0 0.88 878 0.05% max 77 12 11 0.00.0024 182338 100170 3 to 12 15 to 10 4865 13.8

Fig. 17. The top ve bio-diesel producers in 2006.

Source: Ref. [131].

2738

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

others). Bio-diesel fuels have many advantages over petroleum diesel fuel: produce less smoke and particles, have higher cetane number, produce lower carbon monoxide and hydrocarbon emissions, are renewable, biodegradable and non-toxic. When ethyl esters are used as fuel the advantage of totally recyclable carbon dioxide cycle is obtained since ethyl alcohol could be of vegetal origin [3]. The technical disadvantages of bio-diesel/fossil diesel blends include problems with fuel freezing in cold weather, reduced energy density, and degradation of fuel under storage for prolonged periods. One additional problem is encountered when blends are rst introduced into equipment that has a long history of pure hydrocarbon usage. Hydrocarbon fuels typically form a layer of deposits on the inside of tanks, hoses, etc. Bio-diesel blends loosen these deposits, causing them to block fuel lters. However, this is a minor problem, easily remedied by proper lter maintenance during the period following introduction of the bio-diesel blend [136]. Bio-diesel can be blended at any level with petroleum diesel to create a bio-diesel blend. It can be used in compression-ignition (diesel) engines with little or no modications [24]. Several studies show bio-diesel can run in a conventional diesel engine for an extended time. Researchers in several states including Missouri and Idaho, have run diesel engines in pickups, city buses, large trucks and tractors on various mixes of bio-diesel/diesel fuel. These mixtures have ranged from 2/98% (B2), 20/80% (B20) up to 100% (B100) [33,130,137]. Bio-diesel is often used as a blend B20 (20 vol.% bio-diesel and 80 vol.% conventional diesel) rather than using B100 (100 vol.% bio-diesel). It is asserted that 90% of air toxics can be eliminated by using B100 whereas 2040% are reduced using B20. A further assertion is that the risk of illness and life threatening diseases can be reduced using bio-diesel blends [131]. Bio-diesel blends of up to 20% reduce the emissions of HC, CO, SO2, and particulates, as well as improve the engine performance [138,139]. Table 13 shows a summary of engine tests completed at the University of Idaho. These tests were performed with a 100% and a 20% mix of ethyl and methyl ester of rapeseed oil [137].
Table 13 Engine emission results from the University of Idaho Emission Hydrocarbons Carbon monoxide Nitrous oxides Carbon dioxide Particulates Source: Ref. [137]. 100% Ester fuel (B100) (%) 52.4 47.6 10.0 0.9 9.9 20/80 Mix (B20) (%) 19.0 26.1 3.7 0.7 2.8

6. Bio-diesel economy In previous economic studies of bio-diesel production, the main economic criteria were capital cost, manufacturing cost and biodiesel break-even price. The economic performance of a bio-diesel plant (e.g., xed capital cost, total manufacturing cost, and the break-even price of bio-diesel) can be determined once certain factors are identied, such as plant capacity, process technology, raw material cost and chemical costs [13]. At present, the high cost of bio-diesel is the major obstacle to its commercialization. Bio-diesel usually costs over US$0.5/l, compared to US$0.35/l for petroleumbased diesel [84]. The major economic factor to consider for input costs of biodiesel production is the feedstock, which is about 80% of the total operating cost. Other important costs are labor, methanol and catalyst, which must be added to the feedstock [35]. Using an estimated process cost, exclusive of feedstock cost, of US$0.158/l ($0.60/gal) for bio-diesel production, and estimating a feedstock cost of US$0.539/l ($2.04/gal) for rened soy oil, an overall cost of US$0.70/l ($2.64/gal) for the production of soy-based bio-diesel was estimated [117]. Palm oil is the main option that is traded internationally, and with potential for import in the short term [140]. Costs for production from palm oil are estimated; the results are shown in Table 14. Bio-diesel from animal fat is currently the cheapest option ($0.4$0.5/l) while traditional transesterication of vegetable oil is at present around $0.6$0.8/l. Cost reductions of $0.1$0.3/l are expected from economies of scale for new processes. The cost of BTL diesel from ligno-cellulose is more than $0.9/l (feedstock $3.6/GJ), with a potential reduction to $0.7$0.8/l [129]. With pre-tax diesel priced at US$0.18/l in the United States and US$0.200.24/l in some European countries, bio-diesel is thus currently not economically feasible, and more research and technological development will be needed [33,35,49,130,141,142]. The cost of bio-diesel production results in a generally accepted view of the industry in Europe that bio-diesel production is not protable without scal support [142144]. The cost of bio-diesel for IC engine is slightly greater than that of diesel oil (approximate difference of 0.06 4/l). The specic fuel consumption, a function of the engine speed, is higher in bio-diesel than in diesel oil. Bench-test results previously reported showed that the average value of SFC for bio-diesel is 17% greater that that of diesel oil. Other tests on different engines show that the increase ranges from 5% to 20%. Under these conditions the specic cost in 4/kWh shows an increase by about 2.1%. If the cost of the two fuels is the same, the specic cost for the engine running on bio-diesel is 9.8% higher [145].

Table 14 Costs of bio-diesel production Plant size ktpa Million liters 0.33 0.15 0.11 0.08 0.06 0.40 0.40 0.40 0.40 0.40 0.05 0.05 0.05 0.05 0.05 0.11 0.10 0.09 0.08 0.06 0.89 0.70 0.64 0.61 0.56 0.12 0.12 0.12 0.12 0.12 0.77 0.58 0.52 0.49 0.44 0.08 0.08 0.09 0.15 0.15 0.85 0.66 0.61 0.64 0.60 Capital costs Feedstock Methanol Other Total Glycerol credit Net Distribution and blending Total

Tallow-based ($/l) 5 6 20 23 40 46 60 69 120 137 Palm oil ($/l) 50 60 60 71 120 143 Source: Ref. [140].

0.09 0.08 0.06

0.73 0.73 0.73

0.05 0.05 0.05

0.08 0.08 0.06

0.95 0.94 0.89

0.12 0.12 0.12

0.83 0.82 0.78

0.04 0.04 0.04

0.88 0.86 0.82

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741

2739

7. Conclusion The problems with substituting triglycerides for diesel fuels are mostly associated with their high viscosities, low volatilities and polyunsaturated character. The viscosity of vegetable oils, when used as diesel fuel, can be reduced in at least four different ways: (1) dilution with hydrocarbons (blending), (2) emulsication, (3) pyrolysis (thermal cracking), and (4) transesterication (alcoholysis). Transesterication is the most common method and leads to monoalkyl esters of vegetable oils and fats, now called bio-diesel when used for fuel purposes. The main factors affecting transesterication are molar ratio of glycerides to alcohol, catalyst, reaction temperature and pressure, reaction time and the contents of free fatty acids and water in oils. The commonly accepted molar ratios of alcohol to glycerides are 6:130:1. Bio-diesel is a cleaner-burning diesel replacement fuel made from natural, renewable sources such as new and used vegetable oils and animal fats. Just like petroleum diesel, bio-diesel operates in compression-ignition engines or Diesel engines. The bio-diesel was characterized by determining its density, viscosity, high heating value, cetane index, cloud and pour points, characteristics of distillation, and ash and combustion points according to ISO norms. Viscosity is the most important property of bio-diesel since it affects the operation of the fuel injection equipment, particularly at low temperatures when the increase in viscosity affects the uidity of the fuel. The major economic factor to consider for input costs of biodiesel production is the feedstock, which is about 80% of the total operating cost. Other important costs are labor, methanol and catalyst, which must be added to the feedstock. Using an estimated process cost, exclusive of feedstock cost, of US$0.158/l ($0.60/gal) for bio-diesel production, and estimating a feedstock cost of US$0.539/l ($2.04/gal) for rened soy oil, an overall cost of US$0.70/l ($2.64/gal) for the production of soy-based bio-diesel was estimated. Acknowledgement The authors gratefully acknowledges the support provided by Sila Science and expresses his gratitude to its manager. References
[1] Wang YD, Al-Shemmeri T, Eames P, McMullan J, Hewitt N, Huang Y, et al. An experimental investigation of the performance and gaseous exhaust emissions of a diesel engine using blends of a vegetable oil. Appl Thermal Eng 2006;26:168491. [2] Dmytryshyn SL, Dalai AK, Chaudhari ST, Mishra HK, Reaney MJ. Synthesis and characterization of vegetable oil derived esters: evaluation for their diesel additive properties. Bioresour Technol 2004;92:5564. [3] Stavarache C, Vinatoru M, Nishimura R, Maed Y. Fatty acids methyl esters from vegetable oil by means of ultrasonic energy. Ultrason Sonochem 2005;12:36772. [4] Kerschbaum S, Rinke G. Measurement of the temperature dependent viscosity of biodiesel fuels. Fuel 2004;83:28791. [5] Meher LC, Sagar DV, Naik SN. Technical aspects of biodiesel production by transesterication a review. Renew Sustain Energy Rev 2006;10:24868. [6] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol 1999;70:115. [7] Friedrich MS. A worlwide review of the commercial production of biodiesel a technological, economic and ecological investigation based on case studies. Department of Technology and Sustainable Product Management (ITNP), Wien, August 26, 2004. <itnp.wu-wien.ac.at/archiv_doc/BAND41_ FRIEDRICH.PDF>. [8] Dufey A. Biofuels production, trade and sustainable development: emerging issues. Environment Report, International Institute for Environment and Development (IIED), London, UK, November 22, 2006. <www.iied.org/pubs>. [9] Marchetti JM, Miguel VU, Errazu AF. Possible methods for biodiesel production. Renew Sustain Energy Rev 2005;11:130011. [10] Royon D, Daz M, Ellenrieder G, Locatelli S. Enzymatic production of biodiesel from cotton seed oil using t-butanol as a solvent. Bioresour Technol 2007;98:64853.

[11] Watanabe Y, Shimada Y, Sugihara A, Tominaga T. Conversion of degummed soybean oil to biodiesel fuel with immobilized Candida antarctica lipase. J Mol Catal B: Enzym 2002;17:1515. [12] Noureddini H, Zhu D. Kinetics of transesterication of soybean oil. J Am Oil Chem Soc 1997;74:145763. [13] Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste cooking oil: 2. Economic assessment and sensitivity analysis. Bioresour Technol 2003;90:22940. [14] Kusdiana D, Saka S. Kinetics of transesterication in rapeseed oil to biodiesel fuels as treated in supercritical methanol. Fuel 2001;80:6938. [15] Komers K, Stloukal R, Machek J, Skopal F. Biodiesel from rapeseed oil, methanol and KOH: 3. Analysis of composition of actual reaction mixture. Eur J Lipid Sci Technol 2001;103:36371. [16] Acaroglu M, Oguz H, gt H. An investigation of the use of rapeseed oil in agricultural tractors as engine oil. Energy Sour 2001;23:82330. [17] Harrington KJ, DArcy-Evans C. Transesterication in situ of sunower seed oil. Ing Eng Chem Prod Res Dev 1985;24:3148. [18] Peterson CL, Feldman M, Korus R, Auld DL. Batch type transesterication process for winter rape oil. Appl Eng Agric 1991;7:7116. [19] Saifuddin N, Chua KH. Production of ethyl ester (biodiesel) from used frying oil: optimization of transesterication process using microwave irradiation. Malay J Chem 2004;6:7782. [20] Tomasevic AV, Siler-Marinkovic SS. Methanolysis of used frying oil. Fuel Proc Tech 2002;80:16. [21] Demirbas A. Biodiesel from sunower oil in supercritical methanol with calcium oxide. Energy Conv Mgmt 2006;48:93741. [22] Encinar JM, Gonzalez JF, Rodriguez JJ, Tejedor A. Biodiesel fuels from vegetable oils: transesterication of Cynara cardunculus L. oils with ethanol. Energy Fuel 2002;16:44350. [23] Warabi Y, Kusdiana D, Saka S. Reactivity of triglycerides and fatty acids of rapeseed oil in supercritical alcohols. Bioresour Technol 2004;91: 2837. [24] Kim HJ, Kang BS, Kim MJ, Park YM, Kim DK, Lee JS, et al. Transesterication of vegetable oil to biodiesel using heterogeneous base catalyst. Catal Today 2004;9395:31520. [25] Bouaid A, Diaz Y, Martinez M, Aracil J. Pilot plant studies of biodiesel production using Brassica carinata as raw material. Catal Today 2005;106: 1936. [26] Al-Widyan MI, Al-Shyoukh AO. Experimental evaluation of the transesterication of waste palm oil into biodiesel. Bioresour Technol 2002;85:2536. [27] Abigor RD, Uadia PO, Foglia TA, Haas MJ, Jones KC, Okpefa E, et al. Lipasecatalysed production of biodiesel fuel from some Nigerian lauric oils. Biochem Soc Trans 2000;28:97981. [28] Watanabe Y, Shimada Y, Sugihara A. Continuous production of biodiesel fuel from vegetable oil using immobilized Candida antarctica lipase. J Am Oil Chem Soc 2000;77:35560. [29] Watanabe Y, Shimada Y, Sugihara A, Tominaga T. Conversion of degummed soybean oil to biodiesel fuel with immobilized Candida antarctica lipase. J Mol Catal B: Enzym 2002;17:1515. [30] Barnwal BK, Sharma MP. Prospects of biodiesel production from vegetable oils in India. Renew Sust Energy Rev 2005;9:36378. [31] Srivastava A, Prasad R. Triglycerides-based diesel fuels. Renew Sust Energy Rev 2000;4:11133. [32] Demirbas A, Kara H. New options for conversion of vegetable oils to alternative fuels. Energy Sour 2006;28:61926. [33] Balat M. Production of biodiesel from vegetable oils: a survey. Energy Sour A 2007;29:895913. [34] Tippayawong N, Wongsiriamnuay T, Jompakdee W. Performance and emissions of a small agricultural diesel engine fueled with 100% vegetable oil: effects of fuel type and elevated inlet temperature. Asian J Energy Environ 2002;3:13958. [35] Demirbas A. Biodiesel production via non-catalytic SCF method and biodiesel fuel characteristics. Energy Conv Mgmt 2006;47:227182. [36] Ryan TW, Dodge LG, Callahan TJ. The effects of vegetable oil properties on injection and combustion in two different diesel engines. J Am Oil Chem Soc 1984;60:16109. [37] etin M, Yksel F. The use of hazelnut oil as a fuel in pre-chamber diesel engine. Appl Thermal Eng 2007;27:637. [38] Cigizoglu KB, Ozaktas T, Karaosmanoglu F. Used sunower oil as an alternative fuel for diesel engines. Energy Sour 1997;19:55966. [39] Nwafor OMI, Rice G. Performance of rapeseed oil blends in a diesel engine. Appl Energy 1996;54:34554. [40] Labeckas G, Slavinskas S. Performance of direct-injection off-road diesel engine on rapeseed oil. Renew Energy 2006;31:84963. [41] Altn R, etinkaya S, Ycesu HS. The potential of using vegetable oil fuels as fuel for diesel engines. Energy Conv Mgmt 2001;42:52938. [42] Rakopoulos CD, Antonopoulos KA, Rakopoulos DC, Kakaras EC, Pariotis EG. Characteristics of the performance and emissions of a HSDI diesel engine running with cottonseed oil or its methyl ester and their blends with diesel fuel. Int J Vehicle Des 2007;45:20021. [43] Zaher FA. Utilization of used frying oil as diesel engine fuel. Energy Sour 2003;25:81926. [44] Huzayyin AS, Bawady AH, Rady MA, Dawood A. Experimental evaluation of Diesel engine performance and emission using blends of jojoba oil and Diesel fuel. Energy Conv Mgmt 2004;45:2093112.

2740

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741 [81] Lopez DE, Goodwin Jr JG, Bruce DA, Lotero E. Transesterication of triacetin with methanol on solid acid and base catalysts. Appl Catal A: Gen 2005;295:97105. [82] Lee Y, Park SH, Lim IT, Han K, Lee SY. Preparation of alkyl (R)-(2)-3hydroxybutyrate by acidic alcoholysis of poly-(R)-(2)-3-hydroxybutyrate. Enzyme Microb Technol 2000;27:336. [83] Stern R, Hillion G. Purication of esters. Eur Pat Appl EP 356317. (Cl. C07C67/ 56) [c.f. Chem Abstr 113: 58504k; 1990]. [84] Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste cooking oil: 1. Process design and technological assessment. Bioresour Technol 2003;89:116. [85] Reyes-Duarte D, Lopez-Cortes N, Ferrer MJ, Plou F, Ballesteros A. Parameters affecting productivity in the lipase-catalysed synthesis of sucrose palmitate. Biocatal Biotrans 2005;23:1927. [86] Linko YY, Lamsa M, Wu X, Uosukainen W, Sappala J, Linko P. Biodegradable products by lipase biocatalysis. J Biotechnol 1998;66:4150. [87] Ghanem A. The utility of cyclodextrins in lipase-catalyzed transesterication in organic solvents: enhanced reaction rate and enantioselectivity. Org Biomol Chem 2003;1:128291. [88] Shah S, Gupta MN. Lipase catalyzed preparation of biodiesel from Jatropha oil in a solvent free system. Process Biochem 2007;42:40914. [89] De A, Vieira AP, Da Silva MAP, Langone MAP. Biodiesel production via esterication reactions catalyzed by lipase. Lat Am Appl Res 2006;36:2838. [90] Bernardes OL, Bevilaqua JV, Leal MCMR, Freire DMG, Langone MAP. Biodiesel fuel production by the transesterication reaction of soybean oil using immobilized lipase. Appl Biochem Biotechnol 2007;137140:10514. [91] Lai OM, Ghazali HM, Chong CL. Use of enzymatic transesteried palm stearin sunower oil blends in the preparation of table margarine formulation. Food Chem 1999;64:838. [92] Skagerlind P, Jansson M, Bergenstahl B, Hult K. Binding of Rhizomucor miehei lipase to emulsion interfaces and its interference with surfactants. Colloids Surf B: Bioint 1995;4:12935. [93] Stamenkovic OS, Lazic ML, Veljkovic VB, Skala DU. Biodiesel production by enzyme-catalyzed transesterication. Hemijska Ind 2005;59:4959. [94] Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G. Biodiesel analytical methods: August 2002January 2004. Colorado: National Renewable Energy Laboratory (NREL); July 2004 [NREL/SR-510-36240]. [95] Saka S, Kusdiana D. Biodiesel fuel from rapeseed oil as prepared in supercritical methanol. Fuel 2001;80:22531. [96] Kusdiana D, Saka S. Effects of water on biodiesel fuel production by supercritical methanol treatment. Bioresour Technol 2004;91:28995. [97] Balat M. Biodiesel from vegetable oils via transesterication in supercritical ethanol. Energy Edu Sci Technol 2005;16:4552. [98] Vera CR, DIppolito SA, Pieck CL, Parera JM. Production of biodiesel by a twostep supercritical reaction process with adsorption rening. In: 2nd Mercosur congress on chemical engineering and 4th Mercosur congress on process systems engineering (ENPROMER-2005), Rio de Brasil. August 1418, 2005. [99] Demirbas D, Karsloglu S. Biodiesel production facilities from vegetable oils and animal fats. Energy Sour A 2007;29:13341. [100] May CY. Transesterication of palm oil: effect of reaction parameters. J Oil Palm Res 2004;16:111. [101] Bala BK. Studies on biodiesels from transformation of vegetable oils for Diesel engines. Energy Edu Sci Technol 2005;15:143. [102] Demirbas A. Biodiesel production from vegetable oils by supercritical methanol. J Sci Ind Res 2005;64:85865. [103] Ma F, Clements LD, Hanna MA. The effect of catalyst, free fatty acids, and water on transesterication o beef tallow. Trans ASAE 1998;41:12614. [104] Noureddini H, Harkey D, Medikonduru V. A continuous process for the conversion of vegetable oils into methyl esters of fatty acids. JAOCS 1998;75:177583. [105] Feuge RO, Grose T. Modication of vegetable oils. VII. Alkali catalyzed interesterication of peanut oil with ethanol. JAOCS 1949;26:97102. [106] Freeman B, Buttereld RO, Pryde EH. Transesterication kinetics of soybean oil. JAOCS 1986;63:137580. [107] Gupta PK, Kumar R, Panesar PS, Thapar VK. Parametric studies on bio-diesel prepared from rice bran oil. Agric Eng Int: CIGR EJ 2007;IX(April) [Manuscript EE 06 007]. [108] Lee H, Kim SJ, Ahn BS, Lee WK, Kim HS. Role of sulfonic acids in the Sncatalyzed transesterication of dimethyl carbonate with phenol. Catal Today 2003;87:13944. [109] Canakci M, Van Gerpen J. Biodiesel production via acid catalysis. Trans ASAE 1999;42:120310. [110] Meneghetti PSM, Meneghetti MR, Wolf CR, Silva EC, Lima GES, Coimbra DA, et al. Ethanolysis of castor and cottonseed oil: a systematic study using classical catalysts. JAOCS 2006;83:81922. [111] Pinto AC, Guarieiro LLN, Rezende MJC, Ribeiro NM, Torres EA, Lopes WA, et al. Biodiesel: an overview. J Braz Chem Soc 2005;16:131330. [112] Marinetti GV. Hydrolysis of lecithin with sodium methoxide. Biochemistry 1962;1:3503. [113] Marinetti GV. Low temperature partial alcoholysis of triglycerides. J Lipid Res 1966;7:7868. [114] Graboski MS, McCormick RL. Combustion of fat and vegetable oil derived fuels in diesel engines. Prog Energy Combust Sci 1998;24:12564. [115] Darnoko1 D, Cheryan M. Kinetics of palm oil transesterication in a batch reactor. JAOCS 2000;77:12637.

[45] Pramanik K. Properties and use of Jatropha curcas oil and diesel fuel blends in compression ignition engine. Renew Energy 2003;28:23948. [46] Knothe G. Historical perspectives on vegetable oil-based diesel fuels. Inform 2001;12:11037. [47] Ozaktas T. Compression ignition engine fuel properties of a used sunower oildiesel fuel blend. Energy Sour 2000;22:37782. [48] Knothe G, Dunn RO, Bagby MO. Biodiesel: the use of vegetable oils and their derivatives as alternative diesel fuels. In: ACS symposium series no. 666: fuels and chemicals from biomass, Washington, DC, USA. 1997. p. 172208. [49] Demirbas A. Biodiesel fuels from vegetable oils via catalytic and non-catalytic supercritical alcohol transesterications and other methods: a survey. Energy Conv Mgmt 2003;44:2093109. [50] Karaosmonoglu F. Vegetable oil fuels: a review. Energy Sour 1999;21:22131. [51] Fukuda H, Konda A, Noda N. Biodiesel fuel production by transestirication of oils. J Biosci Bioeng 2001;92:40516. [52] Bagby MO. Vegetable oils for diesel fuel: opportunities for development. In: International winter meeting of the ASAE. Chicago: Hyatt Regency; December 1518, 1987. [53] Schwab AW, Bagby MO, Freedman B. Preparation and properties of diesel fuels from vegetable oils. Fuel 1987;66:13728. [54] Lima DG, Soares VCD, Ribeiro BE, Carvalho DA, Cardoso ECV, Rassi FC, et al. Diesel-like fuel obtained by pyrolysis of vegetable oils. J Anal Appl Pyrol 2004;71:98796. [55] Demirbas A. Diesel fuel from vegetable oil via transesterication and soap pyrolysis. Energy Sour 2002;24:83541. [56] Schwab AW, Dykstra GJ, Selke E, Sorenson SC, Pryde EH. Diesel fuel from thermal decomposition of soybean oil. J Am Oil Chem Soc 1988;65:17816. [57] Da Rocha Filho GN, Brodzki D, Djega-Mariadassou G. Formation of alkanes, alkylcycloalkanes and alkylbenzenes during the catalytic hydrocracking of vegetable oils. Fuel 1993;72:5439. [58] Billaud F, Dominguez V, Broutin P, Busson C. Production of hydrocarbons by pyrolysis of methyl esters from rapeseed oil. J Am Oil Chem Soc 1995;72:114954. [59] Alencar JW, Alves PB, Craveiro AA. Pyrolysis of tropical vegetable oils. J Agric Food Chem 1983;31:126870. [60] Akdeniz F, Kck MM, Demirbas A. Liquids from olive husk by using supercritical uid extraction and thermochemical methods. Energy Edu Sci Technol 1998;2:1722. [61] Chand CC, Wan SW. Chinas motor fuels from tung oil. Ind Eng Chem 1947;39:15438. [62] anakc M, zsezen AN. Evaluating waste cooking oils as alternative diesel fuel. GU J Sci 2005;18:8191. [63] Lucia LA, Argyropoulos DS, Adamopoulos L, Gaspar AR. Chemicals and energy from biomass. Can J Chem 2006;8:96070. [64] Vicente G, Martnez M, Aracil J. Integrated biodiesel production: a comparison of different homogeneous catalysts systems. Bioresour Technol 2004;92:297305. [65] Macedo CCS, Abreu FR, Tavares AP, Alves MP, Zara LF, Rubim JC, et al. New heterogeneous metal-oxides based catalyst for vegetable oil transesterication. J Braz Chem Soc 2006;17:12916. [66] Freedman B, Buttereld RO, Pryde EH. Transesterication kinetics of soybean oil. J Am Oil Chem Soc 1986;63:137580. [67] Meher LC, Kulkarni MG, Dalai AK, Naik SN. Transesterication of karanja (Pongamia pinnata) oil by solid basic catalysts. Eur J Lipid Sci Technol 2006;108:38997. [68] Ohi A, Aoyama H, Ohuchi H, Kato A, Yamaoka M. Fatty acid ester from palm oil as diesel fuel. Nenryo Kyokaishi 1983;62:2431. [69] Aksoy HA, Becerik I, Karaosmanoglu F, Yamaz HC, Civelekoglu H. Utilization prospects of Turkish raisin seed oil as an alternative engine fuel. Fuel 1990;69:6003. [70] Gryglewicz S. Rapeseed oil methyl esters preparation using heterogeneous catalysts. Bioresour Technol 1999;70:24953. [71] Korytkowska A, Barszczewska-Rybarek I, Gibas M. Side-reactions in the transesterication of oligoethylene glycols by methacrylates. Des Monom Polym 2001;4:2737. [72] Varghaa V, Truterb P. Biodegradable polymers by reactive blending transesterication of thermoplastic starch with poly(vinyl acetate) and poly(vinyl acetate-co-butyl acrylate. Eur Polymer J 2005;41:71526. [73] Lotero E, Goodwin JG, Bruce DA, Suwannakarn K, Liu Y, Lopez DE. The catalysis of biodiesel synthesis. Catalysis 2006;19:4183. [74] Furuta S, Matsuhasbi H, Arata K. Biodiesel fuel production with solid superacid catalysis in xed bed reactor under atmospheric pressure. Catal Commun 2004;5:7213. [75] Canakci M, Gerpen JV. A pilot plant to produce biodiesel from high free fatty acid feedstocks. Trans ASAE 2003;46:94555. [76] Demirbas A. Biodiesel from vegetable oils via transesterication in supercritical methanol. Energy Conv Mgmt 2002;43:234956. [77] Schuchardta U, Serchelia R, Vargas RM. Transesterication of vegetable oils: a review. J Braz Chem Soc 1998;9:199210. [78] Freier M. Biodiesel alternative fuels offer promise for achieving energy independence. In: Medicinal chemistry symposium. September 2005. <www.nesacs.org/TheNucleus/Sep05>. [79] Goff MJ, Bauer NS, Lopes S, Sutterlin WR, Suppes GJ. Acid-catalyzed alcoholysis of soybean oil. J Am Oil Chem Soc 2004;81:41520. [80] Liu Y, Lotero E, Goodwin Jr JG. Effect of water on sulfuric acid catalyzed esterication. J Molec Catal A: Chem 2006;245:13240.

M. Balat, H. Balat / Energy Conversion and Management 49 (2008) 27272741 [116] Iso M, Chen B, Eguchi M, Kudo T, Shrestha S. Production of biodiesel fuel from triglycerides and alcohol using immobilized lipase. J Molec Catal B: Enzym 2001;16:538. [117] Haas MJ, McAloon AJ, Yee WJ, Foglia TA. A process model to estimate biodiesel production costs. Bioresour Technol 2006;97:6718. [118] Goering E, Schwab W, Daugherty J, Pryde H, Heakin J. Fuel properties of eleven vegetable oils. Trans ASAE 1982;25:147283. [119] Pryor RW, Hanna MA, Schinstock JL, Bashford LL. Soybean oil fuel in a small diesel engine. Trans ASAE 1982;26:3338. [120] Knothe GH. Analytical methods used the production and fuel quality assessment of biodiesel. Trans ASAE 2001;44:193200. [121] Kojima M, Johnson T. Potential for biofuels for transport in developing countries. Energy Sector Management Assistance Programme (ESMAP). Washington (DC): Energy and Water Department, The World Bank Group; October 2005. [122] Collins K. Economic issues related to biofuels. United States Department of Agriculture (USDA) The Ofce of the Chief Economist (OCE), Washington (DC); August 26, 2006. <www.usda.gov/oce/index.htm>. [123] Gubler R. Biodiesel. CEH Marketing Research Report. Chemical Economics Handbook. SRI Consulting; November 2006. 548 p. [124] Boyle G. An overview of alternative transport fuels in developing countries: drivers, status, and factors inuencing market deployment. Hydrogen fuel cells and alternatives in the transport sector: issues for developing countries, United Nations university international conference, Maastricht, The Netherlands, November 79, 2005. [125] European Biodiesel Board (EBB). EBB publishes annual biodiesel production statistics. Press Release, 535/COM/07, Bruxelles, July 17, 2006. [126] Schnepf R. European Union biofuels policy and agriculture: an overview. CRS Report for Congress. March 16, 2006. [127] European Biodiesel Board (EBB). EBB publishes annual biodiesel production statistics. Press Release, 164/COM/06, Bruxelles, April 25, 2006. [128] Sims RH, Hastings A, Schlamadinger B, Taylor G, Smith P. Energy crops: current status and future prospects. Glob Change Biol 2006;12:205476. [129] International Energy Agency (IEA). Biodiesel statistics. IEA energy technology essentials. Paris: OECD/IEA; January 2007. [130] Balat M. Fuel characteristics and the use of biodiesel as a transportation fuel. Energy Sour 2006;28:85564.

2741

[131] Joshi RM, Pegg MJ. Flow properties of biodiesel fuel blends at low temperatures. Fuel 2007;86:14351. [132] Balat M. Current alternative engine fuels. Energy Sour 2005;27: 56977. [133] Demirbas A. Global biofuel strategies. Energy Edu Sci Technol 2006;17: 3363. [134] Prakash CB. A critical review of biodiesel as a transportation fuel in Canada. A technical report. Canada: GCSI Global Change Strategies International, Inc., Transportation Systems Branch; March 25, 1998. [135] Demirbas A. Recent developments in biodiesel fuels. Int J Green Energy 2007;4:1526. [136] Wardle DA. Global sale of green air travel supported using biodiesel. Renew Sust Energy Rev 2003;7:164. [137] Hofman V. Biodiesel fuel. North Dakota: North Dakota State University of Agriculture, Applied Science and US Department of Agriculture Cooperative; 2003. [138] Malhotra RK, Das LM. Biofuels as blending components for motor gasoline and diesel fuels. J Sci Ind Res 2003;62:906. [139] Sastry GSR, Krishna Murthy ASR, Ravi Prasad P, Bhuvaneswari K, Ravi PV. Identication and determination of bio-diesel in diesel. Energy Sour A 2006;28:133742. [140] Dene T, Hole J. Enabling biofuels: biofuel economics. Final Report. Auckland, New Zealand: Minister of Transport, Covec Ltd., June 2006. [141] Bender M. Economic feasibility review for community-scale farmer cooperatives for biodiesel. Biores Technol 1999;70:817. [142] Demirbas MF, Balat M. Recent advances on the production and utilization trends of bio-fuels: a global perspective. Energy Conv Mgmt 2006;47: 237181. [143] United States Department of Agricultural (USDA). EU: Biodiesel industry expanding use of oilseeds. Washington (DC): US Department of Agriculture; September 20, 2003. [144] Balat M. An overview of biofuels and policies in the European Union countries. Energy Sour B 2007;2:16781. [145] Carraretto C, Macor A, Mirandola A, Stoppato A, Tonon S. Biodiesel as alternative fuel: experimental analysis and energetic evaluations. Energy 2004;29:2195211.

You might also like