Vous êtes sur la page 1sur 19

ARTICLE IN PRESS

Available online at www.sciencedirect.com

Fuel xxx (2008) xxxxxx www.fuelrst.com

Review article

Advancements in development and characterization of biodiesel: A review


Y.C. Sharma a,*, B. Singh a, S.N. Upadhyay b
a

Environmental Engineering and Research Laboratories, Department of Applied Chemistry, Institute of Technology, Banaras Hindu University, Varanasi, UP 221 005, India b Department of Chemical Engineering and Technology, Institute of Technology, Banaras Hindu University, Varanasi, UP 221 005, India Received 7 November 2007; received in revised form 21 January 2008; accepted 22 January 2008 Available online 20 February 2008

Abstract An ever increasing demand of fuels has been a challenge for todays scientic workers. The fossil fuel resources are dwindling day by day. Biodiesel seems to be a solution for future. Biodiesel is an environmentally viable fuel. Out of the four ways viz. direct use and blending, micro-emulsions, thermal cracking and transesterication, most commonly used method is transesterication of vegetable oils, fats, waste oils, etc. Latest aspects of development of biodiesel have been discussed in this work. Yield of biodiesel is aected by molar ratio, moisture and water content, reaction temperature, stirring, specic gravity, etc. Biodegradability, kinetics involved in the process of biodiesel production, and its stability have been critically reviewed. Emissions and performance of biodiesel has also been reported. 2008 Elsevier Ltd. All rights reserved.
Keywords: Biodiesel; Transesterication; Catalyst; Stability; Biodegradation

Contents 1. 2. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. Various raw materials used as feedstock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Production of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Acid esterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Alkaline transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A special reference to karanja . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Effect of different parameters on production of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Effect of molar ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Effect of moisture and water content on the yield of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Effect of free fatty acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4. Effect of temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.5. Effect of stirring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6. Effect of specific gravity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Biodegradability of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Stability of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kinetics of the reaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Diesel engine emissions and performance of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 00 00 00 00 00 00 00 00 00 00 00 00 00 00 00 00 00

3. 4.

5. 6. 7. 8.

Corresponding author. Tel.: +91 542 2307025 (O), +91 9935616119 (M); fax: +91 542 2316428. E-mail address: y_sharma_2002@redimail.com (Y.C. Sharma).

0016-2361/$ - see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.fuel.2008.01.014

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
2 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

9. 10. 11. 12.

Indian scenario . . . . . . . . . . . . . . . . . . . . . . . . Cost of biodiesel . . . . . . . . . . . . . . . . . . . . . . . Instrumentation involved in biodiesel production Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

00 00 00 00 00

1. Introduction Whenever I think of fuel, a picture of villages of developing nations eclipsed in darkness, full of poverty, no planned mechanization, agriculture depending on rain God for irrigation comes before my eyes. Be it lighting the house, agriculture, or cooking, each process requires energy. Energy, most countries do not have energy even to light the cities! Superconductivity and hydrogen energy reected some promise but these were completely shortlived. Biodiesel, yes, probably biodiesel is the FUEL of future. Biodiesel is derived from vegetable oils and hence is a renewable fuel. Gasoline and diesel come in the category of non-renewable fuel and will last for a limited period of time. These non-renewable fuels also emit pollutants in the form of oxides of nitrogen, oxides of sulphur, carbon dioxide, carbon monoxide, lead, hydrocarbons, etc. during their processing and use. A renewable fuel such as biodiesel, with lesser exhaust emissions, is the need of the day. Hence, researchers and scientic community worldwide have focused on development of biodiesel and the optimization of the processes to meet the standards and specications needed for the fuel to be used commercially without compromising on the durability of engine parts. The interest in the use of renewable fuel started with the direct use of vegetable oils as a substitute for diesel. However, their direct use in compression ignition engines was restricted due to high viscosity which resulted in poor fuel atomization, incomplete combustion and carbon deposition on the injector and the valve seats causing serious engine fouling [1]. Other constraints of the direct application of vegetable oil were its low volatility and polyunsaturated character. To overcome these constraints, the processes like pyrolysis, micro-emulsication, transesterication, etc. were especially developed. Pyrolysis of the vegetable oil resulted in products with low viscosity, high cetane number, accepted amounts of sulphur, water and sediments, accepted copper corrosion values but were unacceptable in the terms of their ash contents, carbon residues, and pour points. Similarly, micro-emulsion of vegetable oil lowered the viscosity of the oil but resulted in irregular injector needle sticking, heavy carbon deposits and incomplete combustion during 200 h laboratory screening endurance test [2]. Transesterication is a chemical reaction between triglyceride and alcohol in the presence of a catalyst. It consists of a sequence of three consecutive reversible reactions where triglycerides are converted to diglycerides, diglycerides are converted to monoglycerides followed by

the conversion of monoglycerides to glycerol. In each step an ester is produced and thus three ester molecules are produced from one molecule of triglyceride [3]. Out of these three methods, transesterication is the most viable process adopted known so far for the lowering of viscosity. It also gives glycerol as a by-product which has a commercial value. Stoichiometrically, three moles of alcohol are required for each mole of triglyceride, but in general, a higher molar ratio is often employed for maximum ester production depending upon the type of feedstock, amount of catalyst, temperature, etc. Commonly used alcohols include methanol, ethanol, propanol and butanol. However, the yield of biodiesel is independent of the type of the alcohol used and the selection of one of these depends on cost and performance. Methanol is preferred over others due to its low cost [1]. The conventional catalysts used are acid and alkali catalysts depending upon the nature of the oil used for biodiesel production. Another catalyst being studied is lipase. Lipase has advantage over acid and alkali catalysts but its cost is a limiting factor for its use in large scale production of biodiesel. Choice of acid and alkali catalysts depends on the free fatty acids (FFA) content in the raw oil. FFA should not exceed a certain amount for transesterication to occur by an alkali catalyst. Invariably, on all aspects of development of biodiesel, Ma and Hanna [4] have done signicant work. Canakci and Van Gerpan [5,6] reported that transesterication was not feasible if FFA content in the oil was about 3%. Ramadhas et al. [1] and Veljkovic et al. [7] used rubber seed oil and tobacco seed oil, respectively, with higher free fatty acid content (17%). The authors reduced the FFA value to more than 2.0%, which corresponds to 4.0 mg KOH/g, by acid esterication using H2SO4 as a catalyst. Sahoo et al. [8] used zero catalyzed transesterication (using toluene) and acid esterication (using H2SO4) prior to alkaline esterication to reduce the acid value from 22.0% to 2.0%. Sharma and Singh [3] also favored acid esterication prior to alkaline transesterication with karanja oil as feedstock having FFA of 2.53% (5.06 mg KOH/g) using H2SO4. In the same manner, the acid value of jatropha which corresponds to 14% FFA was reduced to less than 1% by using H2SO4 [9]. Table 1 depicts the values of initial FFA of the feedstock, the level reached after acid esterication and the amount of H2SO4 used. After treatment with acid catalyst, H2SO4, the free fatty acid (FFA) value is reduced to less than 2.0% to make transesterication reaction feasible. Table 2 depicts the yield/conversion of biodiesel with dierent oils taken. The yield of biodiesel ranged from 56% from Chlorella

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx Table 1 Values of initial FFA content of dierent feedstock Feedstock Rubber oil Karanja oil Tobacco oil Polanga oil Jatropha oil Mahua oil Karanja oil Initial FFA (%) 17.0 2.53 35.0 22.0 14.0 19.0 2.53 37.96 0.018 40.0 FFA after treatment (%) <2.0 0.95 <2.0 <2.0 <1.0 <1.0 0.3 1.05 0.018 <1.0 Amount (%) and catalyst used 0.5, H2SO4 0.5, H2SO4 1.0/2.0, H2SO4 0.65, H2SO4 1.43, H2SO4 1.0, H2SO4 KOH (appropriate amount) 2.0 Fe2SO4 Reference [1] [3] [7] [8] [9] [10] [11] [16] [18] 3

Table 2 Yield of biodiesel with dierent feedstock Oil taken for study Karanja (Pongamia pinnata) Tobacco (Nicotina tabacum) Polanga (Calophyllum inophyllum) Jatropha (Jatropha curcas) Mahua (Madhuca indica) Karanja (Pongamia pinnata) Karanja (Pongamia pinnata) Soybean (Glycine max) Waste cooking oil Canola oil (Brassica napus) Used frying oil Sunower oil (Helianthus annuus) Chlorella protothecoides Chlorella protothecoides Yield (%) 89.5 91 99 98 9798 97.02 90.04 87.5 56 Conversion (%) 85 92/95 98.4 98 94 Nearly complete >80 Reference [3] [7] [8] [9] [10] [11] [14] [15] [16] [19] [19] [20] [21] [22]

protothecoides to 99% from Jatropha curcas. The conversion of biodiesel ranged from more than 80% from C. prothecoides to 98.4% from soybean oil. Various oils have been in use in dierent countries as raw materials for biodiesel production owing to its availability. Soybean oil is commonly used in United States and rapeseed oil is used in many European countries for biodiesel production, whereas, coconut oil and palm oils are used in Malaysia for biodiesel production [1013]. Transesterication of edible oils has also been carried out from the oil of canola and sunower. Other edible and non-edible oils, animal fats, algae and waste cooking oils have also been investigated by researchers for the development of biodiesel [1431]. Table 3 depicts the work carried out for biodiesel production from various feedstocks under dierent conditions. Type and amount of variables such as feedstock, alcohol, molar ratio, catalyst, reaction temperature, time duration, rate and mode of stirring aects the yield and conversion of biodiesel.

1.1. Various raw materials used as feedstock Leung and Guo [19] compared the transesterication reaction conditions of neat canola oil with used frying oil (UFO). A comparatively higher temperature (333 K),

higher molar ratio (7:1, methanol/UFO), and more amount of catalyst (1.1 wt% NaOH) was needed when compared to edible canola oil where optimal conditions were 315 318 K, 6:1 methanol/oil molar ratio and 1.0 wt% NaOH. However, comparatively a less time (20 min) was needed in UFO for completion of reaction in comparison to canola oil where reaction time took 60 min. Non-edible oils used for transesterication mostly are the oils with higher free fatty acids such as rubber (Ficus elastica), jatropha (J. curcas), karanja (Pongamia pinnata), mahua (Madhuca indica), polanga (Calophyllum inophyllum), tobacco (Nicotina tabacum), etc. In an attempt to reduce the cost of biodiesel, microalgal oils have also been tried by researchers as a source of feedstock for the production of biodiesel due to their higher photosynthetic eciency, higher biomass production and faster growth as compared to other energy crops [3235]. Miao and Wu [21] reported production of biodiesel from microalga C. protothecoides using 100% catalyst quantity (based on oil weight) with 56:1 molar ratio of methanol to oil at temperature of 303 K in 4 h of reaction time. The specic gravity of biodiesel produced reduced from initial value of 0.91 to a nal value of 0.86. Xu et al. [22] obtained high quality and low cost biodiesel from microalgae C. protothecoides of heating value 41 MJ kg1, density 0.864 kg L1 and viscosity 5.2 104 Pa S at 313 K. C. protothecoides have earlier been reported as a feedstock for aquaculture feeds, human food supplements and pharmaceuticals [3639]. Attributing to its low cost, waste cooking oil has also been tried by researchers to develop biodiesel. The idea comes from the fact that triglycerides comprise of greases and oil. Oils are generally in liquid state at room temperature, whereas greases and fats are in solid state at room temperature. Recycled grease is termed as waste grease and is classied as yellow and brown grease depending on free fatty acid composition [17]. The price of yellow grease (FFA < 15%) ranges from $0.04 to $0.09 kg1 while the price of brown grease (FFA > 15%) ranges between 0.004 and $0.014 kg1 [40]. Biodiesel development from grease can reduce its production cost. In United States, an estimate reveals that biodiesel production from 5.2 billion kg/year of greases and animal fats could replace 1.5 million gallons of diesel fuel. The major constraint of direct application of waste frying oil lies in its higher amount of

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

Table 3 Biodiesel production from dierent feedstock Year Ref Feedstock Transesterication stages Alcohol Molar ratio (methanol to oil) 40:1 Catalyst Reaction temperature (K) 473673 (pressure 200 bar) 318 Duration Stirring Conversion/yield Reference Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

ARTICLE IN PRESS

2004

Sunower oil

Single step

Supercritical methanol Supercritical ethanol Methanol Ethanol Methanol

No catalyst

10 40 min

5:1

2005

Pongamia pinnata

Single step

10:1

2005

Madhuca indica

Two step

Methanol

2005

Rubber seed oil

Two step

Methanol

0.300.35 v/v 0.25 v/v 6:1 9:1 56:1 56:1 6:1 7:1

Supercritical CO2 + lipase (Novozym 435) 30 wt% of oil KOH (1% by wt) ZnO Hb-Zeolite Montmorillonite 1% v/v H2SO4 0.7% w/v KOH H2SO4 0.5% by volume NaOH 0.5% by volume Acid catalyst H2SO4 (100%) on the basis of oil weight NaOH 1.0 wt% NaOH 1.1 wt% H2SO4 (1% with lower molar ratio) (2% with higher molar ratio) KOH (1% based on the oil wt.)

6h

7896% conversion with increase in temperature 23% conversion 27% conversion 92% 83% 59% 47% 98% conversion conversion conversion conversion yield

[20]

378 393 333 318 5 318 5 303 303 318 333 333.0 0.1

1.5 h 24 h 1h 1h 20 30 min 30 min

[14]

Magnetic stirrer

[10]

[1]

2006 2006 2006

Chlorella protothecoides Chlorella protothecoides Neat canola oil Used frying oil

Methanol Methanol Single step Methanol

4h 15 min 20 min 25 min

Magnetic stirring 1100 rpm in the rst stage (10 min) and 600 rpm in second stage Magnetic stirrer 400 rpm

More than 80% conversion 63% yield Ester content 98 wt% Ester content 94.6 wt% Yield 91% in 30 min

[22] [21] [19]

2006

Nicotiana tabacum L. (tobacco)

Two step

Methanol

18:1

[7]

6:1

30 min

2006 2006

Pongamia pinnata Soybean oil

Single step

Methanol Methanol

6:1 12:1 4.5:1

KOH (1% by weight) TiO2/ZrO2 (11 wt% Ti) Al2O3/ZrO2 (2.6 wt% Al) K2O/ZrO2 (3.3 wt% K) 1 ml KOH (0.5 M) 1 ml BF3 (12.5%, v/v) NaOH

338 448

2h 1h 2h

Mechanical stirrer 360 rpnm

Yield 9798% Conversion over 95% Conversion 100% Yield 92.54 2.00% Yield 100%

[11] [23]

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

2006

2006

Monosodium glutamate wastewater Soybean oil

Two step

Methanol

0.5 M 12.5%, v/v 6:1

333 353 318

10 min 5 min 10 20 min

[24]

Single step alkali catalyzed

Methanol

2007

2007

Jatropha, pongamia, sunower, soybean, palm Calophyllum inophyllum

Single step

Methanol

3:1

NaOH/KOH (1 wt% of oil)

24 h

Mechanical stirrer 900 rpm Power ultrasonic (frequency 19.7 KHz, power 150 W) Hydrodynamic cavitation (operation pressure 0.7 MPa) Stirring

[25]

[12]

2007

Jatropha curcas

Three step zero catalysed Acid catalysed Alkali catalysed Two step acid catalysed Alkali catalysed Single step

Methanol

6:1

Anhydrous H2SO4 (98.4%) 0.65% by volume KOH 1.5% by weight H2SO4 1.43% v/v (3.5 + acid value, w/v KOH) Anion exchange resin Cation exchange resin (heterogeneous catalyst) Activated CaO (1 wt%) Fe2SO4 KOH H2SO4 NaOH/KOH

338

2h 4h 4h 88 min

Mechanical stirring 450 rpm

Methanol

9:1 0.28 v/v

333

85% yield in 90 min Complete in 4 h reaction >99% yield

[8] Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

ARTICLE IN PRESS

[9]

2007

Triolein

Ethanol

0.16 v/v 10:1

323

24 min 60 min

98.8% purity

[26]

2007 2007 2007

Sunower oil Waste cooking oil Karanja

Single step Two step Two step

Methanol Methanol Methanol

13:1 10:1 6:1 8:1 9:1

333 368 338 318 2

100 min 4h 1h 30 min 30 min

Helix stirrer 1000 rpm No stirring because boiling was sucient Magnetic/mechanical

97.02% conversion 89.5% yield with mechanical 85% yield with magnetic

[27] [16] [3]

ARTICLE IN PRESS
6 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

Fig. 1. Basic scheme for biodiesel production.

FFA. The FFA level of fresh soybean oil has been reported to change from 0.04% to 1.51% after 70 h of frying at 463 K [41]. Hence, due to higher FFA, direct conversion of waste restaurant oil and animal fats to biodiesel via alkaline catalyst is not possible. The level of FFA, therefore, is reduced by an acid catalyst. Following this advantage, researchers have advocated the use of rendered animal fats and restaurant waste oils as biodiesel feedstocks [18,4247]. A major limiting factor of biodiesel is its inverse relationship between its oxidation stability and its cold ow properties. Saturated compounds have good oxidation stability but poor cold temperature properties. Unsaturated compounds have better low temperature properties but fail in oxidation stability [12,17]. Here, the waste oil product can be of advantage over the neat vegetable oil as they have a higher proportion of saturated fatty acids [17] and hence can provide better oxidative stability. Wang et al. [16] have achieved 97.02% yield of biodiesel from waste cooking oil of high acid value i.e. 75.92 0.036 mg KOH/g, by a two step catalysis process. A novel method has been developed by Xue et al. [24] for the production of biodiesel from monosodium glutamate wastewater having COD of 10,000 mg/L. After treatment the COD removal was 85% with 10% formation of crude lipid. The crude lipid was biosynthesized by rhodotorula glutinis by transesterication reaction with methyl ester yield of 92.54 2.0%. Biodiesel has also been synthesized using triolein as a feedstock [26]. Fig. 1 depicts the basic scheme for biodiesel production [48]. 2. Production of biodiesel 2.1. Acid esterication The acid value is a measure of the number of acidic functional groups in a sample and is measured in terms of the quantity of potassium hydroxide required to neutralize the sample. Acid value of the feedstock for alkaline transesterication has to be reduced to less than 2.0 mg KOH/g [15,4951]. However, other authors advocate it to

be less than 4.0 mg KOH/g [1,3]. The commonly used catalyst during acid esterication of neat oil is sulphuric acid (H2SO4) [1,3,710]. Nevertheless, this process too has a drawback as water is produced along with ester from the reaction of FFA with alcohol which inhibits the transesterication of glycerides [17]. For waste cooking oil too the acid employed is sulphuric acid [2931]. But in this case, the conversion reported is low (82%) and the alcohol required for the reaction is high (200% excess of ethanol) [16]. Wang et al. [16] have, therefore, tried a new catalyst Fe2(SO4)3 (ferric sulphate) as an alternate to sulphuric acid and have reported much better conversion (97.02%). As, ferric sulphate is insoluble in oil, it was centrifuged from the liquid after acid esterication and reused for the next batch. High temperature and high concentration of H2SO4 as catalyst could burn some of the oil which will then cause low yield of biodiesel product. 2.2. Alkaline transesterication For oil samples with FFA below 2.0%, alkaline transesterication is preferred over the acid catalyzed transesterication as the former is reported to proceed about 4000 times faster than the latter [2]. The common catalyst employed during alkaline transesterication at industrial level application includes the homogeneous catalysts sodium hydroxide, potassium hydroxide, etc. The use of homogeneous catalyst such as sodium hydroxide and potassium hydroxide has been successful at industrial level for production of biodiesel. However, the biodiesel and glycerin produced have to be puried to remove the basic catalyst and need its separation by washing with hot distilled water twice or thrice. Thus, heterogeneous catalyst has also been tried by researchers to overcome this drawback of time consumption and colossal consumption of water. The heterogeneous catalyst can be separated from the nal product by ltration which checks time consumption and prevents the consumption of large volume of water. The ltered solid then can be reused. The application of a heterogeneous catalyst, CaO has been

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx 7

tested by Grandos et al. for its feasibility [27]. The experiments conrmed that CaO could be used as a catalyat for the transesterication reaction without signicant deactivation up to eight runs with signicant amount of CaO. Leung and Guo [19] tried three dierent homogeneous catalysts i.e. sodium hydroxide (NaOH), potassium hydroxide (KOH), and sodium methoxide (CH3ONa). The optimum requirement of the catalyst were 1.1, 1.3 and 1.5 wt% for NaOH, CH3ONa, and KOH respectively for the maximum ester content. The amount of NaOH required was less than the amounts of both the CH3ONa or KOH for the same conversion of fatty acid methyl ester as NaOH has lower molar mass (40 g/mol), compared to CH3ONa (54 g/mol) and KOH (56 g/mol). However, in terms of yield, CH3ONa proved to be a better catalyst than NaOH and KOH because CH3ONa dissociates into CH3 O and Na+ and does not form any water as side product. On the other hand, NaOH and KOH forms sodium (or potassium) methoxide when dissolved in methanol and produces water. This water then reacts with Na+ (or K+) to form sodium (or potassium) soaps. Potassium methoxide (KOCH3) has also been used as an alkali catalyst by dissolving potassium hydroxide in methanol [52]. Sharma and Singh [3] reported better yield with NaOH as a catalyst over KOH while using magnetic stirrer. Whereas, when mechanical stirrer was adopted the yield was same with equal amount of NaOH and KOH (0.5 wt%). However, during the separation of the nal products from glycerol, KOH was more convenient. Potassium soaps being softer than sodium soaps did not block the bottom of separating funnel unlike latter and were removed easily. Hence, KOH as catalyst is preferred over NaOH at industrial level application [3,19]. The base catalyzed reaction is reported to be very sensitive to the purity of the reactant. FFA content should not exceed beyond a certain limit. The eciency of the reaction was aected to some extent when FFA content exceeded 0.5 wt% [28]. However, when the feedstock is waste cooking oil, the limit of FFA is a bit relaxed and FFA content a little beyond 1.0 wt% did not have any eect on the methyl ester conversion [16,53]. The amount of catalyst (KOH) required was 1.0 wt% to reach 97.02% conversion of biodiesel. Ramadhas et al. [1] have reduced the acid value to less than 2.0% through acid catalyst followed by alkaline transesterication. The amount of catalyst used for alkaline transesterication ranged between 0.3% and 1.0%. The maximum conversion eciency was reported at 0.5% of NaOH during alkaline transesterication. Excess amount of catalyst gave rise to formation of an emulsion. This increased the viscosity and led to the formation of gels. Conversion eciency decreased to 60% when the catalyst amount was increased to 0.8% (wt of NaOH/wt of oil). However, esterication also did not take place without sucient amount of the catalyst. Srivastava and Verma [54] used sodium methoxide solution prepared by dissolving NaOH (28.5 g) and methanol

(2.0 kg) as an alkaline catalyst for transesterication. Sahoo et al. [8] used alkaline transesterication after reducing the free fatty acid value from 44 mg KOH/g (i.e. 22%) to less than 4 mg KOH/g (i.e. <2%) through zero catalyzed and acid catalyzed transesterication. 1.5% by weight of potassium hydroxide was found sucient for the maximum yield of ester. Sarin et al. [12] prepared a series of biodiesel from the edible oils such as sunower, soybean, palm and non-edible oils such as jatropha and karanja. The catalyst used was either NaOH or KOH with 1 wt % of oil. The amount of alkali catalyst was calculated on the basis of the amount needed to neutralize the unreacted acids plus 0.35% for virgin oil which came out to be 0.55% w/v KOH [9]. 1.0% KOH was reported as the optimal amount for alkaline transesterication reaction of karanja oil by Meher et al. [11] and Karmee and Chadha [14]. Karmee and Chadha also tried Hb-Zeolite, montmorillonite K-10 and ZnO but the conversion of fatty acid methyl ester was less. Around 83% conversion with ZnO was possible only with a longer reaction time of 24 h. Even lesser conversions of 59% and 47% were achieved with HbZeolite and montmorillonite K-10, respectively, in comparison to 92% achieved by KOH. While using the rened karanja oil, the conversion reached to 99% with 0.5% of NaOH or CH3ONa [14]. 1% KOH was optimum amount of catalyst even with used frying oil reported by Marinkovic and Tomasevic [55]. Ghadge and Raheman [10] calculated the amount of catalyst to be 0.7% w/v KOH as the cumulative sum of 0.5% for catalyst plus the amount needed to neutralize the unreacted acids (2 mg KOH/g) i.e. 0.2%. NaOH was used as a catalyst by Ji et al. [25] with power ultrasonic and hydrodynamic cavitation methods and found the methods superior as compared to mechanical stirring. The amount of KOH used for the transesterication of tobacco (Nicotiana tabacum) oil after acid esterication as reported by Veljkovic et al. [7] is 1.0% (based on oil weight) with a yield of 91%. Using the same oil Usta got a yield of 86%. Hence, the two step process of acid and alkaline esterication can be expected to yield a better result when compared to one step base catalyzed transesterication [56]. In an attempt to nd a new catalyst, Furuta et al. [23] tested amorphous zirconia solid catalysts, TiO2/ZrO2 (11 wt% Ti) and Al2O3/ZrO2 (2.6 wt% of Al) and reported more than 95% conversion. The reason for this is the amphoteric nature of zirconia. However, the temperature required was quite high i.e. 448473 K. The homogeneous catalysts require neutralization and separation steps from the reaction mixture. To accomplish this, water, solvents and energy are needed. To overcome these, heterogeneous based catalysts have been tested by researchers, where separation is possible without using solvent and shows easy regeneration. The nature of the nal product has been less corrosive in character and the whole process is termed safer, cheaper and more environment friendly [57]. The heterogeneous catalyst being used by researchers include alkalineearth oxides, zeolites,

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
8 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

hydrotalcites, MgO and CaO [27,5762]. In addition, reactor heating was not needed when MgO was used as a catalyst in batch process [57]. Another method where neither acid nor base catalyst is required is non-catalytic transesterication where supercritical methanol is used instead of methanol and nearly complete conversion is achieved. The reason is that, the oil and supercritical methanol exist in the single phase [20,6367]. The advantage with supercritical methanol is that the conversion gets 95% complete in 10 min. But at the same time, higher molar ratio (40:1) is employed. The supercritical condition is based on the eect of relationship between pressure and temperature upon the thermophysical properties of the solvent, such as dielectric constant, viscosity, specic weight and polarity. There is a decrease in the dielectric constant of methanol in supercritical state. Fig. 2 depicts the inuence of temperature and alcohol used in the synthesis of biodiesel. In methanol, conversion increased from 78% to 96% with increase in temperature. The conversion was even higher with supercritical ethanol

as the solubility parameter of ethanol is closer to the solubility parameter of the oil. But, high temperature and high pressure conditions i.e. 573 K and 20 MPa are needed for the reaction to proceed and complete. [13,20]. Another added advantage with heterogeneous based catalyst is the lesser consumption of catalyst. As per the study by Zhang et al. [29] annually 88 tonnes of sodium hydroxide is needed for 8000 tonnes of biodiesel production. While as per the simulation study by Dossin et al. [57] the requirement of MgO will be only 5.7 tonnes for 100,000 tonnes production of biodiesel. 3. A special reference to karanja Among the several indigenous plant species, karanja is one of the medium sized plants found in several parts of India. The plant is also said to be highly tolerant to salinity and is reported to be grown in various soil textures viz. stony, sandy and clayey. Karanja can grow in humid as well as subtropical environments with annual rainfall ranging between 500 and 2500 mm. This is one of the reasons for wide availability of this plant species. The constituents of karanja oil viz. furanoavones, furanoavonols, chromenoavones, avones and furanodiketones make the oil non-edible and hence the oil is underutilized. The presence of these constituents, however, gives the oil antifungal characteristics and enhances its application in medicinal ointments. The woody part of the tree is still in use by tribal people where the plant is locally available such as brushing of teeth. The oil expelled from the seeds is also burned during the festival of lighting to purify the environment. All these applications are at local or regional level and 94% of the oil from plant is still underutilized. The present production of karanja oil approximately is 200 million tons per annum [11]. However, its production potential is much more i.e. 135,000 million tones per annum [8]. The time needed by the tree to mature ranges from 4 to 7 years and depending on the size of the tree the yield of kernels per tree is between 8 and 24 kg. The oil content extracted by various authors ranges between 30.0 to 33% [3,14]. The oil is used by common people due to its low cost and easy availability. The fatty acid composition of karanja oil has been reported in Table 4. Karanja oil comprises of
Table 4 Dierent fatty acids present in karanja oil [107] Fatty acid (common name) Palmitic Stearic Oleic Linoleic Arachidic Gadoleic Behenic Lignoceric Systematic name Formula Structure wt%

Fig. 2. Synthesis of biodiesel at various temperatures in supercritical (a) methanol, (b) ethanol (j) 473 K, () 523 K, (N) 573 K, (.) 623 K, () 673 K.

Hexadecanoic Octadecanoic cis-9-octadecenoic cis-9, cis-12-octadecadienoic Eicosanoic 11-eicosenoic Docosanoic Tetracosanoic

C16H32O2 C18H36O2 C18H34O2 C18H32O2 C20H40O2 C20H38O2 C22H44O2 C24H48O2

16:0 18:0 18:1 18:2 20:0 20:1 22:0 24:0

10.6 6.8 49.4 19.0 4.1 2.4 5.3 2.4

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx Table 5 Test method and limits of biodiesel along with its comparison with karanja oil methyl ester [12] Property (units) Flash point (C) Viscosity at 40 C (cSt) Sulphated ash (mass%) Sulphur (mass%) Cloud point (C) Copper corrosion Cetane number Water and sediment (vol. %) CCR 100% (mass%) Neutralization value (mg, KOH/gm) Free glycerin (mass%) Total glycerin (mass%) Phosphorus (mass%) Distillation temperature Oxidation stability, h ASTM 6751 test method D-93 D-445 D-874 D-5453 D-2500 D-130 D-613 D-2709 D-4530 D-664 D-6584 D-6584 D-4951 D-1160 NA ASTM 6751 limits Min. 130 1.96.0 Max. 0.02 Max. 0.05 NA Max. 3 Min. 47 Max. 0.05 Max. 0.05 Max. 0.80 Max. 0.02 Max. 0.24 Max. 0.001 90% at 360 C NA IS 15607 test method IS 1448 P:21 IS 1448 P:25 IS 1448 P:4 ASTM D 5453 IS 1448 P:10 IS 1448 P:15 IS 1448 P:9 D-2709 D-4530 IS 1448 P:1/Sec. 1 D-6584 D-6584 D-4951 Not under spec. EN 14112 IS 15607 limits Min. 120 2.56.0 Max. 0.02 Max. 0.005 NA Max. 1 Min. 51 Max. 0.05 Max. 0.05 Max. 0.50 Max. 0.02 Max. 0.25 Max. 0.001 Min. 6 h Karanja oil methyl ester 141 4.16 0.002 0.003 4 1 55.1 0.03 <0.01 0.10 0.01 0.01 <0.001 90% 2.35 9

29.2% saturated fatty acids and 70.8% unsaturated fatty acids. The maximum proportion comprises of oleic acid (cis-9-octadecenoic) i.e. 49.4%, whereas, gadoleic acid (11-eicosenoic, 2.4%) and linoceric acid (tetracosanoic, 2.4%) are in traces. Table 5 depicts the ASTM 6751 and IS 15607 test methods and limits for biodiesel. In respect of karanja, it is observed that it meets all the specication of American Society for Testing and Materials (ASTM) and Indian Standards but fails in oxidation stability test. Hence, appropriate methods have to be devised to make the karanja oil methyl ester t for commercialization. For the transesterication reaction, Sharma and Singh [3] advocated for 8:1 molar ratio during acidic esterication so as to reduce the acid value below 2 mg KOH/g i.e. 1.0% FFA. For the alkaline transesterication reaction, the authors advocate 9:1 molar ratio for the completion of the reaction. Mechanical stirrer produced better results than magnetic stirrer and resulted in higher yield (89.5%) in 1 h with 0.5% H2SO4 and 0.5% NaOH at 318 2 K. Meher et al. [11] have neutralized the oil by using potassium hydroxide to reduce the acid value of the oil from 5.06 mg KOH/g to 0.6 mg KOH/g. The yield of 97.0 98.0% was achieved in 2 h with 6:1 molar ratio (methanol:oil); and in 1 h with 12:1 molar ratio. The optimum temperature was 338 K with the rate of stirring 360 rpm and 1% KOH as catalyst. Karmee and Chadha [14] have reported 92% conversion with karanja oil using 10:1 molar ratio (methanol:oil) at 333 K with 1.0% KOH by weight. 4. Eect of dierent parameters on production of biodiesel 4.1. Eect of molar ratio Ramadhas et al. [1] and Sahoo et al. [8] reported 6:1 molar ratio during acid esterication and 9:1 molar ratio (alcohol:oil) during alkaline esterication to be the optimum amount for biodiesel production from high FFA rubber seed oil and polanga seed oil respectively. Sharma and

Singh [3] used similar two step transesterication and took 8:1 molar ratio for acid esterication and 9:1 molar ratio for alkaline esterication for optimum yield of biodiesel production from karanja oil. Veljkovic et al. [7] used 18:1 molar ratio during acid esterication and 6:1 molar ratio during alkaline esterication. Meher et al. [11] carried out investigation with 6:1 molar ratio during acid esterication and 12:1 molar ratio during alkaline esterication. Instead of taking molar ratio, Tiwary et al.[9] and Ghadge and Raheman [10] used volume as a measure of ratio. While Tiwary et al. used 0.28 v/v (methanol/oil) during acid esterication and 0.16 v/v (methanol/oil) during alkaline esterication, Ghadge and Raheman used 0.300.35 v/v (methanol/oil) during acid esterication and 0.25 v/v (methanol/oil) during alkaline esterication. Karmee and Chadha [14] used a single step transesterication and have achieved 92% conversion by taking 10:1 molar ratio. Presence of sucient amount of methanol during the transesterication reaction is essential to break the glycerine-fatty acid linkages [68]. But excess of methanol should be avoided. Increasing the molar ratio of methanol/oil beyond 6:1 neither increases the product yield nor the ester content, but rather makes the ester recovery process complicated and raised its cost. Leung and Guo [19] suggested that methanol has polar hydroxyl group which can act as an emulsier causing emulsication. Thus separation of the ester layer from the water layer becomes dicult. Miao and Wu [21] have reported that addition of large quantity of methanol, i.e. 70:1 and 84:1 molar ratio slowed down the separation of the ester and glycerol phases during the production of biodiesel. 56:1 molar ratio was reported to be optimum for transesterication of microalgal oil. 4.2. Eect of moisture and water content on the yield of biodiesel Kusdiana and Saka [69] observed that water could pose a greater negative eect than presence of free fatty

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
10 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

Fig. 3. Plots for yields of methyl esters as a function of water content in transesterication of triglycerides.

Fig. 4. Eect of FFA on the yield of methyl ester during alkali catalyzed transesterication.

4.4. Eect of temperature acids and hence the feedstock should be water free. Romano [70] and Canakci and Van Gerpen [5] insisted that even a small amount of water (0.1%) in the transesterication reaction would decrease the ester conversion from vegetable oil. Demirbas [13] too reported a decrease in yield of the alkyl ester due to presence of water and FFA as they cause soap formation, consume catalyst and reduce the eectiveness of catalyst. Srivastava and Verma [54] removed the moisture content from the vegetable oil by heating in oven for 1 h at 383 K. Meher et al. [11] too reported a precautionary step to prevent moisture absorbance and maintenance of catalytic activity by preparing the fresh solution of potassium hydroxide and methanol. Ellis et al. [52] found that even a small amount of water in the feedstock or from esterication reaction producing water from FFA might cause reduction in conversion of fatty acid methyl ester and formation of soap instead. At the same time the presence of water had a positive eect in the yield of methyl esters when methanol at room temperature was substituted by supercritical methanol. However, no explanation for this has been provided. [13]. Fig. 3 depicts the inuence of water content on yield of methyl esters. It is observed that acid catalyst is most prone to presence of water followed by alkaline catalyst. No eect on ester content was observed when supercritical methanol was used. The presence of water had negligible eect on the conversion while using lipase as a catalyst [20,71]. 4.3. Eect of free fatty acids Free fatty acids (FFAs) content after acid esterication should be minimal or otherwise less than 2% FFAs. These FFAs react with the alkaline catalyst to produce soaps instead of esters. Fig. 4 depicts the eect of FFAs on the yield of methyl ester during alkali catalysed transesterication. There is a signicant drop in the ester conversion when the free fatty acids are beyond 2% [72]. The temperature maintained by the researchers during dierent steps range between 318 and 338 K. The boiling point of methanol is 333.7 K. Temperature higher than this will burn the alcohol and will result in much lesser yield. A study by Leung and Guo [19] showed that temperature higher than 323 K had a negative impact on the product yield for neat oil, but had a positive eect for waste oil with higher viscosities. 4.5. Eect of stirring Stirring can play an important role in the yield of biodiesel production. Meher et al. [11] conducted the transesterication reaction with 180, 360 and 600 revolutions per minute (rpm) and reported incomplete reaction with 180 rpm. The yield of methyl ester was same with 360

Fig. 5. The variation of product specic gravity with reaction time under dierent molar ratio of methanol to oil. Reaction conditions: 303 K, 160 rpm, 100% catalyst quantity based on oil weight.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx Table 6 Comparison of dierent technologies to produce biodiesel Variable Reaction temperature (K) Free fatty acid in raw materials Water in raw materials Yield of methyl esters Recovery of glycerol Purication of methyl esters Production cost of catalyst Alkali catalysis 333343 Sapanoed products Interference with reaction Normal Dicult Repeated washing Cheap Lipase catalysis 303313 Methyl esters No inuence Higher Easy None Relatively expensive Supercritical alcohol 512658 Esters Good Medium Acid catalysis 328353 Esters Interference with reaction Normal Dicult Repeated washing Cheap 11

and 600 rpm. Sharma and Singh [3] reported that mode of stirring too plays a vital role in the transesterication reaction. The yield of biodiesel increased from 85% to 89.5% when magnetic stirrer (1000 rpm) was replaced with mechanical stirrer (1100 rpm). A plausible explanation may be a thorough mixing of the reactants by mechanical stirrer. 4.6. Eect of specic gravity Lower value of the specic gravity of the nal product is an indication of completion of reaction and removal of heavy glycerine. The inuence of molar ratio, temperature and catalyst quantity on the specic gravity of the biodiesel was studied by Miao and Wu [21]. The specic gravity of the product decreased sharply up to 2 h of reaction time using 30:1 molar ratio and up to 4 h of reaction time using 45:1 and 56:1 molar ratio after which it was almost constant. The best process combination reduced the product specic gravity from 0.912 to 0.864 with 100% catalyst, 56:1 molar ratio at 303 K in 4 h of reaction time. Fig. 5 depicts the change in specic gravity with reaction time under dierent molar ratio of methanol to oil. A comparison of dierent technologies to produce biodiesel is shown in Table 6 [48]. 5. Biodegradability of biodiesel Biodiesel is reported to be highly biodegradable in freshwater as well as soil environments. 9098% of biodiesel is mineralized in 2128 days under aerobic as well as anaerobic conditions [7375]. Biodiesel has been reported to remove twice the amount of crude oil from sand as conventional shoreline cleaners [76]. Biodiesel increases the biodegradability of crude oil by means of cometabolism. More

than 98% degradation of pure biodiesel after 28 days is reported by Pasqualino et al. [73] in comparison to 50% and 56% by diesel fuel and gasoline respectively. Also, the time taken to reach 50% biodegradation reduced from 28 to 22 days in 5% biodiesel mixture and from 28 to 16 days in case of 20% biodiesel mixture at room temperature. The biodegradability of the mixture was reported to increase with addition of biodiesel. Table 7 depicts the biodegradability of fossil diesel under dierent conditions [77 80]. 6. Stability of biodiesel Biodiesel, chemically is fatty acid methyl ester if alcohol used during transesterication is methanol or fatty acid ethyl ester in case of ethanol. This ester molecule will get hydrolyzed to alcohol and acid in the presence of air. Conversion of ester into alcohol will lead to reduction in ash point whereas conversion of ester into acid will increase the total acid number. This makes the biodiesel unstable on storage [12]. The stability of biodiesel also depends on the feedstock used for the biodiesel production. The feedstock with larger proportion of saturated fatty acids will be more stable than those having larger proportion of unsaturated fatty acids. But again, higher portion of saturated fatty acid lowers the low temperature properties such as cloud and pour points. Hence, a major drawback of biodiesel lies in its tradeo between the level of saturation of biodiesel and its cold ow properties [12,17]. The oxidation stability of biodiesel is not dependent on the total number of double bonds but on the total number of bis-allylic sites (the methylene CH directly adjacent to the two double bonds). These esters undergo auto-oxidation which is dependent on the number and position of the double bonds and forms by-product such as acids, esters, aldehydes, ketones, lactones, etc. [12,8183]. Fatty acid methyl esters form a radical next to the double bond during the oxidation process. This radical binds with the oxygen in air, which is a biradical to form peroxide radical. A new radical is created from the fatty acid methyl ester by this peroxide radical which binds with oxygen in air. This augments the auto-oxidation cycle at an exponential rapid rate whereby 100 new radicals are created quickly from one single radical resulting in the formation of a series of by-products. The fuel thus gets deteriorated as there is formation of sediment and gum. Peroxide formation

Table 7 Biodegradability of fossil diesel % degradation Articially contaminated soils Using soil columns Fuel contaminated soil in the arctic Under aerobic conditions Under anaerobic conditions 67 81 9095 42 18 No. of days 109 310 365 30 50 Reference [77] [78] [79] [80] [80]

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
12 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

through this route leads to obligomerization even at ambient temperature. As the deterioration of biodiesel is attributed to the formation of peroxide in the initial step, the remedy suggested is to prevent peroxide formation during the stages of biodiesel manufacture and throughout its distribution chain [12,81]. As biodiesels are primarily made from vegetable oils which do contain naturally occurring antioxidants such as tocopherols, sterols and tocotrienols and these remain in the biodiesel during the process of manufacture. However, the distillation and purication step destroys these natural antioxidants and hence becomes prone to oxidation. Synthetic antioxidants such as phenolic types or aminic types have to be added to make it stable and hence acceptable in market [12,81,8385]. Sarin et al. [12] suggested blending jatropha biodiesel with palm biodiesel to reduce the antioxidant dosage by 8090% to maintain good oxidation stability.

converted to glycerol. In each of the steps an ester is formed. Thus three esters are obtained from one triglycerides molecule. TG M $ DG FAME
K4 K1

2 3 4

DG M $ MG FAME
K5

K2

MG M $ G FAME
K6

K3

7. Kinetics of the reaction Transesterication reaction consists of a sequence of reversible reactions [52].

where, TG, triglycerides; DG, diglycerides; MG, monoglycerides; M, methanol and FAME, fatty acid methyl esters. The values of rate constants for forward reactions, K1, K2 and K3, are found to be 5.00, 4.93 and 29.67 dm3 mol1 min1 and the values of the rate constants of backward reactions involved in the kinetics of biodiesel development, K4, K5, and K6 are reported to be 3.54, 2.99 and 0.79 dm3 mol1 min1. A software MATLAB 6.1 was used to generate concentration prole of glycerol, tri-,di-, and monoglycerides and fatty acid methyl ester. Fig. 6 shows the concentration proles of all components in which a plateau is reached at equilibrium [52].

In the rst step, triglycerides are converted to diglycerides which get converted to monoglycerides in the next step. In the third and last step, monoglycerides are

8. Diesel engine emissions and performance of biodiesel Researchers worldwide are on a consensus that biodiesel, irrespective of the feedstock used results in a decrease in the emissions of hydrocarbons (HC), carbon monoxide (CO), particulate matter (PM) emissions and sulphur dioxide (SO2). Only oxides of nitrogen (NOx) are reported to increase which is due to oxygen content in the biodiesel [8694]. It is also said to be carbon neutral as it contributes no net carbon dioxide to the atmosphere [9597]. However, in a study conducted by Sahoo et al. [8] the NOx emission from 100% biodiesel lowered to 4% for polanga seed oil. This is attributed by the authors to dierence in engine geometry, compression ratio, less reaction time and temperature. A remarkable decrease in the emission of unburnt hydrocarbon was observed. 40% reduction of CO2 emissions was observed for B20 and B100 biodiesel. Thermal eciency of the engine also improved by 0.1%. Smoke emissions reduced by 35% in the case of B60 biodiesel. The results were obtained without any engine hardware modications. Other authors have suggested various strategies to eliminate the NOx emission. Szybist et al. [98] suggested to change the chemical composition of feedstock by increasing the methyl oleate and addition of cetane improv-

Fig. 6. Simulated concentration prole of glycerol (G), tri-glyceride (TG), di-glyceride (DG), mono-glyceride (MG) and fatty acid methyl ester (FAME) during transesterication.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx Table 8 Emission impact of 20 vol% biodiesel for soybean based biodiesel added to an average base diesel fuel Percent change in emissions NOx (nitrogen oxides) PM (particulate matter) HC (hydrocarbons) CO (carbon monoxides) +2.0 10.1 21.1 11.0 13

ers. This reduces the iodine value. Boehman et al. [99] and Kegl [100] recommended to retard the injection timing which could lead to decrease in NOx emission. Nabi et al. [101] advocated for the exhaust gas recirculation to reduce the NOx emission. Kegl [102] stressed on the importance of fuel injection system to reduce the engine emissions as well as fuel consumption. Author suggests that pressure squareness (ratio of mean to maximum injection pressure) should be at maximum and fuelling in the rst part of injection has to be less to reduce NOx emission. Simultaneously, the fuelling in the last part of injection should be less to reduce the smoke emissions. Shahid and Jamal [103] reported that blend of biodiesel till B20 (20% biodiesel and 80% diesel) would

be safe above which it would cause maintenance problem and might damage the engine. A study on comparison of carbonyls emissions from diesel and biodiesel blend was carried by Correa and Arbilla [104]. Higher emission of formaldehyde, acetaldehyde, acrolein, acetone, propionaldehyde, and butyraldehyde was observed from B2, B5, B10 and B20 blends than neat diesel. However, signicant reduction in emission in terms of benzaldehyde was observed from biodiesel blends. Srivastava and Verma [54] have reported the HC, CO and NO emissions from karanja oil methyl ester to be slightly higher as compared with petrodiesel. HC emission of diesel at maximum load was 85 ppm, while that of biodiesel was 120 ppm due to poor mixing with air. CO emission of diesel at maximum load was reported to be 0.18% as compared to 0.21% of biodiesel and NO emission was reported to be 12% higher than that of biodiesel. Maximum thermal eciency of methyl ester and brake specic fuel consumption of the biodiesel were quite close to that of diesel and hence the authors were of the view that karanja oil methyl ester can replace diesel as an alternative fuel. Table 8 depicts emission impact of 20 vol% for soybean based biodiesel. It can be seen that there is a signicant reduction

Table 9 List of 26 indigenous plant meeting US, Germany and European standards (a oil from kernel, b oil from seeds, osa: other saturated acids, uk: unknown) [107] Sources Rhus succedanea Linn Annona reticulate Linn Ervatamia coronaria Stapf Thevetia peruviana Merrill Basella rubra Linn Canarium commune Linn Celastrus paniculatus Linn Terminalia bellirica Roxb Vernonia cinerea Less Corylus avellana Jatropa curcas Linn Putranjiva roxburghii Calophyllum apetalum Wild Calophyllum inophyllum Linn Mesua ferrea Linn Azadirachta indica Moringa concanensis Nimmo Moringa oleifera Lam Pongamia pinnata Pierre Ziziphus mauritiana Lam Sapindus trifoliatus Linn Schleichera oleosa Oken Madhuca indica JF Gmel Mimusops hexendra Robx Pterygota alata Rbr Holoptelia integrifolia Oil 39.5a 42.0b 41.6b 67.0a 36.9b 73.0a 52.0a 40.0b 38.0a 57.5a 40.0b 41.8a 47.5a 65.0a 68.5a 44.5a 35.5b 35.0a 33.0b 33.0b 45.5a 40.0b 40.0b 47.0a 35.0b 37.4b SN 204.0 203.6 201.1 201.5 202.9 204.6 236.6 198.8 205.2 200.5 202.6 199.6 200.4 201.4 201.0 201.1 199.7 199.7 196.7 198.6 195.0 193.0 202.1 202.0 202.6 208.7 IV 92.6 87.2 76.0 84.0 85.3 77.3 77.5 77.8 68.5 84.51 93.0 82.9 97.6 71.5 81.3 69.3 76.0 75.4 80.9 81.8 64.5 57.9 74.2 62.2 98.4 49.9 CN 52.22 53.47 56.33 57.48 54.00 55.58 51.9 56.24 57.51 54.50 52.31 54.99 51.57 57.3 55.10 57.83 56.32 56.66 55.84 55.37 59.77 61.55 56.61 59.32 51.09 61.22 Fatty acid composition (%) 16:0 (25.4); 18:1 (46.8); 18:2 (27.8) 14:0 (1.0); 16:0 (17.2); 16:1 (4.2); 18:0 (7.5); 18:1 (48.4); 18:2 (21.7) 16:0 (24.4); 16:1 (0.2); 18:0 (7.2); 18:1 (50.5); 18:2 (15.8); 18:3 (0.6); 20:0 (0.7); 20:1 (0.2); 22:0 (0.2); uk (0.2) 16:0 (15.6); 18:0 (10.5); 8:1 (60.9); 18:2 (5.2); 18:3 (7.4); 20:0 (0.3); 22:0 (0.1) 14:0 (0.4); 16:0 (19.7); 16:1 (0.4); 18:0 (6.5); 18:1 (50.3); 18:2 (21.6); 18:3 (0.4); 20:4 (0.7) 16:0 (29.0); 18:0 (9.7); 18:1 (38.3); 18:2 (21.8); 18:3 (1.2) 1:0 (2.0); 2:0 (1.7); 16:0 (25.1); 18:0 (6.7); 18:1 (46.1); 18:2 (15.4); 18:3 (3.0) 16:0 14:0 14:0 14:0 16:0 16:0 (35.0); 18:1 (24.0); 18:2 (31.0); osa (10.0) (8.0); 16:0 (23.0); 18:0 (8.0); 18:1 (32.0); 18:2 (22.0); 20:0 (3.0); 22:0 (4.0) (3.2); 16:0 (3.1); 18:0 (2.6); 18:1 (88.0); 18:2 (2.9); uk (0.2) (1.4); 16:0 (15.6); 18:0 (9.7); 18:1 (40.8); 18:2 (32.1); 20:0 (0.4) (8.0); 18:0 (15.0); 18:1 (56.0); 18:2 (18.0); 20:0 (3.0) (8.0); 18:0 (14.0); 18:1 (48.0); 18:2 (30.0)

16:0 (17.9); 16:1 (2.5); 18:0 (18.5); 18:1 (42.7); 18:2 (13.7); 18:3 (2.1); 24:0 (2.6) 14:0 (0.9); 16:0 (10.8); 18:0 (12.4); 18:1 (60.0); 18:2 (15.0); 20:0 (0.9) 16:0 (14.9); 18:0 (14.4); 18:1 (61.9); 18:2 (7.5); 20:0 (1.3) 16:0 (9.7); 18:0 (2.4); 18:1 (83.8); 18:2 (0.8); 20:0 (3.3) 16:0 (9.1); 16:1 (2.1); 18:0 (2.7); 18:1 (79.4); 18:2 (0.7); 18:3 (0.2); 20:0 (5.8) 16:0 (10.6); 18:0 (6.8); 18:1 (49.4); 18:2 (19.0); 20:0 (4.1); 20:1 (2.4); 22:0 (5.3); 24:0 (2.4) 16:0 (10.4); 18:0 (5.5); 18:1 (64.4); 18:2 (12.4); 20:0 (1.8); 20:1 (2.6); 22:0 (1.2); 22:1 (1.7) 16:0 (5.4); 18:0 (8.5); 18:1 (55.1); 18:2 (8.2); 20:0 (20.7); 22:0 (2.1) 16:0 (1.6); 16:1 (3.1); 18:0 (10.1); 18:1 (52.5); 20:0 (19.7); 22:0 (4.0); 22:1 (0.9); gadolic acid (8.4) 14:0 (1.0); 16:0 (17.8); 18:0 (14.0); 18:1 (46.3); 18:2 (17.9); 20:0 (3.0) 16:0 (19.0); 18:0 (14.0); 18:1 (63.0); 18:2 (3.0); 20:0 (1.0) 16:0 (14.5); 18:0 (8.5); 18:1 (44.0); 18:2 (32.4); uk (1.0) 14:0 (3.5); 16:0 (35.1); 16:1 (1.9); 18:0 (4.5); 18:1 (53.3); 20:0 (1.1); uk (1.4)

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
14 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

Table 10 List of 11 species meeting the US biodiesel standards on the basis of fatty acid composition, SN, IV and CN (a oil from kernel, b oil from seeds, osa: other saturated acids, uk: unknown) [107] Sources Vallaris solanacea Kuntze Balanites roxburghii Planch Croton tiglium Linn Hydnocarpus wightiana Blume Mappia foetida Milers Perilla frutescens Britton Aphanamixis polystachya Park Princeptia utilis Royle Meyna laxiora Robyns Aegle marmelos correa Roxb Tectona grandis Linn Oil 33.0b 43.0a 45.0b 63.0a 48.0b 40.5b 35.0a 37.2a 38.5b 34.0b 44.5a SN 198.3 188.9 203.9 210.5 200.7 199.0 203.8 201.9 202.8 202.5 200.9 IV 104.7 109.9 102.9 102.1 101.3 193.9 109.1 108.4 101.3 114.9 111.3 CN 50.26 50.46 49.9 49.25 50.70 30.09 48.52 48.94 50.42 48.30 48.31 Fatty acid composition (%) 16:0 (7.2); 18:0 (14.4); 18:1 (35.3); 18:2 (40.4); 20:0 (1.8); 22:0 (0.4); 24:0 (0.5) 16:0 (17.0); 16:1 (4.3); 18:0 (7.8); 18:1 (32.4); 18:2 (31.3); 18:3 (7.2) 14:0 (11.0); 16:0 (1.2); 18:0 (0.5); 18:1 (56.0); 18:2 (29.0); 20:0 (2.3) 16:0 (1.8); 18:1 (6.9); hydnocarpic acid (48.7); gorlic acid (12.2); chaulmoogric acid (27.0); chaulmoogtic homolog (3.4) 16:0 (7.1); 18:0 (17.7); 18:1 (38.4); 18:3 (36.8) 18:1 (9.8); 18:2 (47.5); 18:3 (36.2); osa (6.5) 16:0 (23.1); 18:0 (12.8); 18:1 (21.5); 18:2 (29.0); 18:3 (13.6) 14:0 (1.8); 16:0 (15.2); 18:0 (4.5); 18:1 (32.6); 18:2 (43.6); 24:0 (0.9) uk (1.4) 16:0 (18.8); 18:0 (9.0); 18:1 (32.5); 18:2 (39.7) 16:0 (16.6); 18:0 (8.8); 18:1 (30.5); 18:2 (36.0); 18:3 (8.1) 14:0 (0.2); 16:0 (11.0); 18:0 (10.2); 18:1 (29.5); 18:2 (46.4); 18:3 (0.4); 20:0 (2.3)

in HC (21%), CO (10%), PM (11%) and 2.0% increase in NOx emission [105]. 9. Indian scenario India imports more than 40% of its edible oil requirement and hence non-edible oils are used for the development of biodiesel. India is a agrarian nation and has rich plant biodiversity which can support the development of biodiesel. India also has a vast geographical area with agricultural lands as well as wastelands on which oil bearing plants can be planted. Common non-edible oil bearing plants and trees include neem, karanja, mahua, jatropha, etc. The oil yields from these species at present are insucient to meet the demand for raw material on large scale production of biodiesel. Hence, there has been government initiatives and interest from few private rms to enhance the production and distribution facilities of biodiesel throughout the country. The Petroleum Ministry has set a target for biodiesel to meet 20% of Indias diesel demand. Governments initiative has resulted in large scale plantation of J. curcas in the state Andhra Pradesh. Oil and Natural Gas Corporation (ONGC) has planned to build an export oriented renery at kakinada in Andhra Pradesh which will have a annual production capacity of 5.57.5 million tonnes [106]. Azam et al. [107] have studied the prole of 75 indigenous plant species of India containing 30% or more oil in their seed, fruit or nut. Out of these plants, based on saponication number, iodine value, cetane number and fatty acid composition, 26 species were found to be most suitable for use as biodiesel and they met the biodiesel standards of USA (ASTM D 6751-02, ASTM PS 121-99), Germany (DIN V 51606) and European Standard Organization (EN 14214). Another 11 plant species met the specications of US biodiesel standards. Authors predicts that cultivation of Azardirachta indica or P. pinnata on 40.09 million

and 19.9 million ha, respectively, will meet the target of 100% replacement of imported biodiesel which amounted to 87.5 million tons in 20032004. These 37 species are listed in Tables 9 and 10. Among these species, jatropha (Jatropha curcas), karanja (P. pinnata), neem (A. indica), mahua (Madhuca indica) and polanga (Calophyllum inophyllum) have catched the attention of researchers and biodiesel manufacturer in India and feasibility of rest of the plant species still remains unexplored. 10. Cost of biodiesel Various factors contributing to the cost of biodiesel include raw material, other reactants, nature of purication, its storage, etc. However, the major factor which contributes the cost of biodiesel production is the feedstock, which is about 80% of the total operating cost [108]. Cost of biodiesel reported by Zhang et al. [29] is US $0.5 l1 as compared to US $0.35 for normal diesel. Bender in his review has reported the cost of biodiesel to be US $0.30 l1 and US $0.69 l1 when the fuel was produced from soybean and rapeseed respectively. The intact oilseed was taken as starting material while calculating the cost. Canakci and Van Gerpen while using a small pilot scale plant in a batch process estimated the cost to be US $0.42 l1 while using rened, bleached and deodorized soybean oil. The prots from glycerol and capital cost for operation was not included [109,110]. Haas et al. [111] in a study of review on biodiesel production cost found the feedstock to contribute a substantial portion in production cost. A process model was prepared by the author to estimate biodiesel production costs. Taking all the factors into account viz. raw material (vegetable oil, methanol, catalysts), utilities (electricity, etc.), labour, supplies, general works and depreciation, the cost of biodiesel was estimated to be US $0.561 l1. The coproduct glycerol as 80% w/w aqueous solution was valued to be

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx 15

US $0.034 l1 and reduced the production cost of biodiesel by 6% on deduction of this value from original cost. Hence, the gross operating cost of biodiesel was estimated to be US $0.527 l1, of which the feedstock comprised 87.36% of the overall cost of biodiesel production. Nas and Berktay [112] too concluded that feedstock cost is the major component contributing to the cost of biodiesel production. In a comparison of feedstock viz. soybean oil and yellow grease (recycled cooking oil from restaurants), it was estimated that yellow grease is quite less expensive that soybean oil. On the contrary, the supply of yellow grease is limited and has other applications and hence it cannot be used on a large scale production. An eight year comprehensive study (from 20042005 to 20122013), on the price rise of petroleum diesel, yellow grease and soybean oil is reported. Cost of petroleum has been estimated to rise from 0.67 to 0.75 (i.e. 11.94% increase). Biodiesel fuel obtained from yellow grease is estimated to rise from 1.41 to 1.55 (i.e. 9.93% increase). Biodiesel fuel from soybean oil will rise from 2.54 to 2.80 (i.e. 10.24% increase). Author is of opinion that biodiesel will not be produced at a cost comparable with that of petrodiesel unless the soybean oil price decline.

Biodiesel comprises of 11% oxygen by weight which improve the combustion process and hence has reduced emissions in terms of hydrocarbon, carbon monoxide and particulate emissions but increases nitrogen oxide emissions. Nitrogen oxide emissions from biodiesel can be reduced by adding cetane enhancers that are di-tert-butyl peroxide at 1% or 2-ethylhexyl nitrate at 0.5%. However, addition of these enhancers will increase the cost of biodiesel production. Nevertheless, reduction in emission of oxides of nitrogen is desired as they are ozone precursors and after reaching stratosphere it will result in ozone layer depletion. In Indian context, the government has xed the price of biodiesel delivered at reneries to be Rs. 26.50 l1. This rate of biodiesel has been xed after the Indian government has provided subsidies in order to encourage its application [113]. The cost of biodiesel after blending with petrodiesel will reduce as the cost of biodiesel becomes less signicant in blended form. At present, biodiesel can be blended with 80% petrodiesel (B20) without any engine modication. B 100 (100% biodiesel) costed US $3.76 per gallon of biodiesel in USA in June 2006, whereas, B 20 (20% biodiesel and 80% petrodiesel) costed US $2.98 per gallon of biodiesel.

Table 11 Instrumentation involved in characterization of biodiesel Feedstock Tobacco Polanga seed oil Pongamia pinnata Instrument involved for characterization High performance liquid chromatograph High performance liquid chromatograph (i) HPLC (Perkin-Elmer series 200) equipped with refractive index detector (shodex RI 71) (ii) H NMR Bruker DPX 300 spectrometer (Bruker, Rheinstetten, Germany) (iii) Gas chromatograph (for analysis of fatty acid composition of karanja oil) (i) GC (Model HR/GC/5300) [for fatty acid composition of vegetable oils) (ii) Thin layer chromatography (TLC) Gas chromatography (Nucon, India) equipped with a FID detector Gas chromatography (GC) equipped with a capillary column (SPBTM-5, 30 m 0.32 mm 0.25 lm) and a FID (i) HP 6890 series II Gas Chromatograph with a 3365/II GC-chemstation and a FID (ii) Thin layer chromatography (TLC) (i) Gel permeation chromatograph (GPC) (ii) Gas chromatograph (GC), Nucon 5765, India CE-440 elemental analyzer (for determining elemental composition of biodiesel) Gas Chromatographymass spectrometric analysis (i) X-ray diractometer (ii) GPC with an RI (refractive index) detector, (carrier: THF) (iii) Gas chromatographymass spectrometer Gas chromatograph HPLC system (Hitachi, Ltd, Tokyo, Japan, D-7000 interface, L-7100 For FAME determination Agilent 6890 GC with a HP INNOwax capillary column For catalyst characterization (i) Balzer Prisma quadrupole mass spectrometer (QMS 200) (ii) Seifert 3000 XRD diractometer (equipped with a PW gonimeter) (iii) VG Escalab 200 R spectrometer (equipped with a hemispherical electron analyzer) (iv) Nicolet 5700 Fourier transform spectrophotometer (equipped with an Hg-col-Te cryodetector) Gas chromatographymass spectrometer Gas chromatography (i) Size exclusion chromatography (SEC) (ii) Fourier-transformed infrared spectroscopy Reference [7] [8] [11]

Jatropha, Pongamia, sunower, soybean, palm Pongamia pinnata Waste cooking oil Neat and used frying oil Sunower oil Microalgal oil Microalga chlorella protothecoides Soybean oil

[12] [14] [16] [19] [20] [21] [22] [23]

Lipid extracted from the cell of R. glutinis Triolein Sunower oil

[24] [26] [27]

Biopiles Canola, corn, peanut, olive, waste vegetable Castor oil

[79] [85] [104]

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
16 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx

11. Instrumentation involved in biodiesel production For production of biodiesel, there are some primary steps. If kernel or seeds are taken as starting materials, in rst step the raw oil is to be expelled from the kernel by mechanical crusher. This raw oil is in fact feedstock for development of biodiesel. After feedstock is obtained, beakers, measuring asks, test tubes, separating funnels, etc. are needed up to development of biodiesel. Practically no instruments are needed till its development. Most of the instruments are used for characterization of the biodiesel product. Further, the instruments used for characterization can be divided into the categories of minor and major equipments. Viscometer, cetane number analyzer, etc. are minor equipments. Flash point, cloud point, pour point, etc. are also determined by minor instruments. Major instruments are however, needed for chemical characterization of biodiesel product. Gas chromatograph, gas chromatographmass spectrometer, high performance liquid chromatograph, fourier transformed infrared spectrometer, elemental analyzer etc. are the major instruments used for chemical characterization of the product. The major instruments are listed in Table 11. 12. Conclusions Biodiesel is derived from a varied range of vegetable oil (edible and non-edible), animal fats, used frying oil, waste cooking oil and wastewater. The edible oil in use at present is soybean, sunower, canola, palm. The non-edible oil used as feedstock for biodiesel production includes J. curcas, P. pinnata, M. indica, F. elastica, A. indica, C. inophyllum, etc. The main advantage in its usage is attributed to lesser exhaust emissions in terms of carbon monoxide, hydrocarbons, particulate matter, polycyclic aromatic hydrocarbon compounds and nitrited polycyclic aromatic hydrocarbon compounds. Biodiesel is said to be carbon neutral as more of carbon dioxide is absorbed by the biodiesel yielding plants than what is added to the atmosphere when used as fuel. Exhaust emissions of NOx can be controlled by adopting certain strategies such as change in composition of feedstock, addition of cetane improvers, retardation of injection timing, exhaust gas recirculation, etc. Transesterication is the process successfully employed at present to reduce the viscosity of biodiesel and improve other characteristics. Methanol being cheaper is the commonly used alcohol during transesterication reaction. Among the catalysts, homogeneous catalysts such as sulphuric acid, sodium hydroxide, potassium hydroxide are commonly used at industrial level production of biodiesel. Heterogeneous catalysts such as calcium oxide, magnesium oxide and others are also being tried to decrease the catalyst amount and production cost of biodiesel. Transesterication reaction can be completed even without catalyst by using supercritical methanol but it will increase the production cost of biodiesel as it is energy intensive. The molar ratio of alcohol to oil required is 3:1 by stoichiometry, but

excess molar ratio has been used for biodiesel production for better yield in lesser time. The molar ratio employed during acid esterication is between 6:1 and 18:1 whereas the molar ratio used alkaline transesterication ranges between 5:1 and 12:1 after reducing the acid value to less than 2.0% approximately. The temperature ranges between 318 and 338 K as the boiling point of methanol is 337.7 K and heating beyond this temperature would burn methanol. However, higher temperature is employed while using supercritical methanol (473573 K). Depending on the feedstock taken; amount and type of alcohol and catalyst; temperature employed; mode and rate of stirring; there is dierence in the yield of biodiesel which varied from 80 to 100%. The percent conversion of biodiesel also ranged between 80% and 100%. Another added advantage of biodiesel is that it is biodegradable in nature. When used as blend along with diesel fuel, it shows positive synergic eect of biodegradation by means of cometabolism. Major disadvantage of biodiesel is the inverse relationship of oxidation stability of biodiesel with its low temperature properties which includes cloud point and pour point. Higher composition of saturated fatty acids in feedstock will increase the oxidation stability of biodiesel but will lower its cloud and pour points. Whereas, higher composition of unsaturated fatty acids will enhance the cloud point and pour point of biodiesel but will have a poor oxidation stability. Hence, a balance has to be maintained between the ratio of saturates and unsaturates for the oil to be used as a feedstock for biodiesel production. Edible oils are in use in developed nations such as USA and European nations but developing nations are not self sucient in the production of edible oils and hence have emphasized in the application of a number of non-edible oils. In a country like India, which is rich in plant biodiversity, there are many plant species whose seeds remain unutilized and underutilized have been tried for biodiesel production. These species have shown promises and fullls various biodiesel standards. However, there still is paucity in terms of all the standards which should be fullled for the large commercial application and its acceptance from public and governing bodies. References
[1] Ramadhas AS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA rubber seed oil. Fuel 2005;84:33540. [2] Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterication of oils. J Biosci Bioeng 2001;92:40516. [3] Sharma YC, Singh B. Development of biodiesel from karanja, a tree found in rural India. Fuel 2007. doi:10.1016/j.fuel.2007.08.001. [4] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol 1999;70:115. [5] Canakci M, Gerpan JV. Biodiesel production via acid catalysis. Trans Am Soc Agric Eng 1999;42(5):120310. [6] Canakci M, Gerpan JV. Biodiesel production from oils and fats with high free fatty acids. Trans Am Soc Agric Eng 2001;44:142936. [7] Veljkovic VB, Lakicevic SH, Stamenkovic OS, Todorovic ZB, Lazic ML. Biodiesel production from tobacco (Nicotiana tabacum L.) seed oil with a high content of free fatty acids. Fuel 2006;85:26715.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx [8] Sahoo PK, Das LM, Babu MKG, Naik SN. Biodiesel development from high acid value polanga seed oil and performance evaluation in a CI engine. Fuel 2007;86:44854. [9] Tiwari AK, Kumar A, Raheman H. Biodiesel production from jatropha oil (Jatropha curcas) with high free fatty acids: an optimized process. Biomass Bioenergy 2007;31:56975. [10] Ghadge SV, Raheman H. Biodiesel production from mahua (Madhuca indica) oil having high free fatty acids. Biomass Bioenergy 2005;28:6015. [11] Meher LC, Dharmagadda VSS, Naik SN. Optimization of alkalicatalyzed transesterication of Pongamia pinnata oil for production of biodiesel. Bioresour Technol 2006;97:13927. [12] Sarin R, Sharma M, Sinharay S, Malhotra RK. Jatropha-palm biodiesel blends: an optimum mix for Asia. Fuel 2007;86:136571. [13] Demirbas A. Biodiesel production via non-catalytic SCF method and biodiesel fuel characteristics. Energy Conservat Manage 2006;47:227182. [14] Karmee SK, Chadha A. Preparation of biodiesel from crude oil of Pongamia pinnata. Bioresour Technol 2005;96:14259. [15] Canakci M, Gerpen JV. Biodiesel production via acid catalysis. Trans Am Soc Agric Eng 1999;42:120310. [16] Wang Y, Ou S, Liu P, Zhang Z. Preparation of biodiesel from waste cooking oil via two-step catalyzed process. Energy Conservat Manage 2007;48:1848. [17] Canakci M. The potential of restaurant waste lipids as biodiesel feedstocks. Bioresour Technol 2007;98:18390. [18] Canakci M, Gerpen JV. A pilot plant to produce biodiesel from high free fatty acid feedstocks. Trans Am Soc Agric Eng 2003;46: 94554. [19] Leung DYC, Guo Y. Transesterication of neat and used frying oil:Optimization for biodiesel production. Fuel Process Technol 2006;87:88390. [20] Madras G, Kolluru C, Kumar R. Synthesis of biodiesel in supercritical uids. Fuel 2004;83:202933. [21] Miao X, Wu Q. Biodiesel production from heterotrophic microalgal oil. Bioresour Technol 2006;97:8416. [22] Xu H, Miao X, Wu Q. High quality biodiesel production from a microalga Chlorella protothecoides by heterotrophic growth in fermenters. J Biotechnol 2006;126:499500. [23] Furuta S, Matsuhashi H, Arata K. Biodiesel fuel production with solid amorphous-zirconia catalysis in xed bed reactor. Biomass Bioenergy 2006;30:8703. [24] Xue F, Zhang X, Luo H, Tan T. A new method for preparing raw material for biodiesel production. Process Biochem 2006;41: 1699702. [25] Ji J, Wang J, Li Y, Yu Y, Xu Z. Preparation of biodiesel with the help of ultrasonic and hydrodynamic cavitation. Ultrasonics 2006;44:4114. [26] Kitakawa NS, Honda H, Kuribayashi H, Toda T, Fukumura T, Yonemoto T. Biodiesel production using anionic ion-exchange resin as heterogeneous catalyst. Bioresour Technol 2007;98:41621. [27] Granados ML, Poves MDZ, Alonso DM, Mariscal R, Galisteo FC, Tost RM, et al. Biodiesel from sunower oil by using activated calcium oxide. Appl Catal B: Environ 2007;73:31726. [28] Freedman B, Buttereld RO, Pryde EH. J Am Oil Chem Soc 1986;63:137580. [29] Zhang Y, Dube MA, McLean DDl. Biodiesel production from waste cooking oil: 1. Process design and technological assessment. Bioresour Technol 2003;89:116. [30] Al-Widyan MI, Al-Shyoukh AO. Experimental evaluation of the transesterication of waste palm oil into biodiesel. Bioresour Technol 2002;85:2536. [31] Tashtoush GM, Al-Widyan MI, Al-Shyoukh AO. Experimental study on evaluation and optimization of conversion of waste animal fat into biodiesel. Energy Convers Manage 2004;45:2697711. [32] Milne TA, Evans RJ, Nagle N. Catalytic conversion of microalgae and vegetable oils to premium gasoline, with shape-selective zeolites. Biomass 1990;21:21932. 17

[33] Ginzburg BZ. Liquid fuel (oil) from halophilic algae: a renewable source of non-polluting energy. Renew Energy 1993;3:24952. [34] Dote Y, Sawayama S, Inoue S, Minowa T, Yokoyama S. Recovery of liquid fuel from hydrocarbon-rich microalgae by thermochemical liquefaction. Fuel 1994;73:18557. [35] Minowa T, Yokoyama SY, Kishimoto M, Okakurat T. Oil production from algal cells of Dunaliella tertiolecta by direct thermochemical liquefaction. Fuel 1995;74:17358. [36] Kyle DJ. Production and use of lipids from microalgae. Lipid Technol 1992;4(3):5964. [37] Running JA, Huss RJ, Olson PT. Heterotrophic production of ascorbic acid by microalgae. J Appl Phycol 1994;4:99104. [38] Borowitzka MA. Microalgae source of pharmaceuticals and other biologically active compounds. J Appl Phycol 1995;7:315. [39] Chen F. High cell density culture of microalgae in heterotrophic growth. Trends Biotechnol 1996;14:4216. [40] Simenson J. Personal communication. Simenson Rendering. Quimby, 2000, IA, USA. [41] Tyagi VK, Vasishtha AK. Changes in the characteristics and composition of oils during deep-fat frying. JAOCS 1996;73(4): 499506. [42] Engelman HW, Guenther DA, Silvis TW. Vegetable oil as a diesel fuel. SAE Paper 1978, 78-DGP-19. [43] Lague CM, Lo KV, Staley LM. Waste vegetable oil as a diesel fuel extender. Canadian Agric Eng 1988;30(1):2732. [44] Kouremenos DA, Rakopoulos CD, Kotsiopoulos PN, Yfantis EA. Experimental investigation of waste olive oil utilization as a fuel supplement in a diesel engine. In: Energy and the environment into the 1990s-proceedings of the rst world renewable energy congress, Reading, UK, 1990. p. 211822. [45] Karaosmanoglu F, Isigigur A, Hamdullahpur F, Gulder OL, Aksoy HA. Used Canola oil as a diesel fuel alternative. In: Renewable energy, technology and the environment-proceedings of the second world renewable energy congress, Reading, UK, 1992. p. 14559. [46] Cigizoglu KB, Ozaktas T, Karaosmanoglu F. Used Sunower oil as an alternative fuel for diesel engines. Energy Sources 1997;19(6): 55966. [47] Ozbay N, Oktar N, Tapan NA. Esterication of free fatty acids in waste cooking oils (WCO): role of ion-exchange resins. Fuel 2008. doi:10.1016/j.fuel.2007.12.010. [48] Marchetti JM, Migual VU, Errazu AF. Possible methods for biodiesel production. Renew Sustain Energy Rev 2007;11:130011. [49] Freedman B, Pryde EH. Fatty esters from vegetable oils for use as a diesel fuel. In: Vegetable oils fuels-proceedings of the international conference on plant and vegetable oils as fuels. ASAE Publication 482, Fargo, ND, USA, 1982. p. 11722. [50] Liu K. Preparation of fatty acid methyl esters for gas-chromatographic analysis of lipids in biological materials. J Am Oil Chem Soc 1994;71(11):117987. [51] Mittelbach M, Pokits B, Silberholz A. Production and fuel properties of fatty acid methyl esters from used frying oil. In: Liquid fuels from renewable resources. Proceedings of an alternative energy conference. ASAE Publication Nashville, TN, USA, 1992. p. 748. [52] Ellis N, Guan F, Chen T, Poon C. Monitoring biodiesel production (transesterication) using in situ viscometer. Chem Eng J 2007. doi:10.1016/j.cej.2007.06.034. [53] Cetinkaya M, Karaosmanoglu F. Optimization of base-catalyzed transesterication reaction of used cooking oil. Energy Fuels 2004;18(6):188895. [54] Srivastava PK, Verma M. Methyl ester of karanja oil as an alternative renewable source energy. Fuel 2007. doi:10.1016/ j.fuel.2007.08.018. [55] Marinkovic SS, Tomasevic A. Methanolysis of used frying oil. Fuel Process Technol 2003;81:16. [56] Usta N. Use of tobacco oil methyl ester in turbocharged indirect injection diesel engine. Biomass Bioenergy 2005;28:7786.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
18 Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx [81] McCormick RL, Ratcli M, Moens L, Lawrence R. Several factors aecting the stability of biodiesel in standard accelerated tests. Fuel Process Technol 2007;88:6517. [82] Frankel EN. Lipid oxidation. 2nd ed. Bridgewater: The Oily Press; 2005. [83] Waynick JA. Characterization of biodiesel oxidation and oxidation products: CRC Project no. AVFL-2b. National Renewable Energy Laboratory, NREL/TP-540-39096.2005. [84] Tatandjiiska RB, Marekov IN, NikilovaDamyanova BM. Determination of triacylglycerol classes and molecular species in seed oils with high content of linoleic and linolenic acids. J Sci Food Agric 1996;72:40310. [85] VanGerpen JH, Hammond EG, Yu L, Monyem A. Determining the inuence of contaminants on biodiesel properties, SAE Technical Paper No. 971685, 1997. [86] Dorado MP, Ballesteros E, Gomez JM, Lopez FJ. Exhaust emissions from a diesel engine fueled with transesteried waste olive oil. Fuel 2003;82:13115. [87] Senda J, Okui N, Tsukamoto T, Fujimoto H. On board measurement of engine performance and emissions in diesel vehicle operated with biodiesel fuel. SAE; 2004. 2004-01-0083. [88] Dorado MP, Ballesteros E, Arnal JM, Gomez J, Gimenez FJL. Testing waste olive oil methyl ester as a fuel in a diesel engine. Energy Fuels 2003;17:15605. [89] Usta N. An experimental study on performance and exhaust emissions of a diesel engine fuelled with tobacco seed oil methyl ester. Energy Conservat Manage 2005;46:237386. [90] Labeckas G, Slavinskas S. The eect of rapeseed oil methyl ester on direct injection Diesel engine performance and exhaust emissions. Energy Conservat Manage 2006;47:195467. [91] Sinha S, Agarwal AK. Performance evaluation of a biodiesel fuelled transport diesel engine. SAE 2005; 2005-01-1730. [92] Turrio-Baldassarry L. Emissions comparison of urban bus engine fuelled with diesel oil and biodiesel blend. Sci Total Environ 2004;327:14762. [93] Monyem A, Gerpen JH. The eect of biodiesel oxidation on engine performance and emission. Biomass Energy 2001;20: 31725. [94] Canakci M, Erdil A, Arcaklioglu E. Performance and exhaust emission of a biodiesel engine. Appl Energy 2006;83: 594605. [95] Lang X, Dalai AK, Bakhshi NN, Reaney MJ, Hertz PB. Preparation and characterization of bio-diesels from various bio-oils. Bioresour Technol 2001;80:5362. [96] Antolin G, Tinaut FV, Briceno Y, Castano V, Perez C, Ramirez AL. Optimization of biodiesel production by sunower oil transesterication. Bioresour Technol 2002;83:1114. [97] Vicente G, Martinez M, Aracil J. Integrated biodiesel production: a comparison of dierent homogeneous catalysis systems. Bioresour Technol 2004;92:297305. [98] Szybist JP, Boehman AL, Taylor JD, McCormick RL. Evaluation of formulation strategies to eliminate the biodiesel NOx eect. Fuel Process Technol 2005;86:110926. [99] Boehman AL, Morris D, Szybist J. The impact of the bulk modulus of diesel fuels on fuel injection timing. Energy Fuels 2004;18: 187782. [100] Kegl B. Experimental investigation of optimal timing of the diesel engine injection pump by using biodiesel fuel. Fuels Energy 2006;20:146070. [101] Nabi N, Akhter S, Shahadat MZ. Improvement of engine emissions with conventional diesel fuel and dieselbiodiesel blends. Bioresour Technol 2006;97:3728. [102] Kegl B. Biodiesel usage at low temperature. Fuel 2007. doi:10.1016/ j.fuel.2007.06.023. [103] Shahid EM, Jamal Y. A review of biodiesel as vehicular fuel. Renew Sustain Energy Rev 2007. doi:10.1016/j.rser.2007.06.001. [104] Correa SM, Arbilla G. Carbonly emissions in diesel and biodiesel exhaust. Atmos Environ 2008;42:76975.

[57] Dossin TF, Reyniers MF, Berger RJ, Marin GB. Simulation of heterogeneously MgO-catalyzed transesterication for ne-chemical and biodiesel industrial production. Appl Catal B: Environ 2006;67:13648. [58] Meyer U, Hoelderich WF. Transesterication of methyl benzoate and dimethyl terephthalate with ethylene glycol over basic zeolites. Appl Catal A: General 1999;178:15966. [59] Corma A, Iborra S, Miquel S, Primo J. Catalysis for the production of ne chemicals: production of food emulsiers, monoglycerides, by glycerolyis of fats with solid base catalysis. J Catal 1998;173:31521. [60] Grylewicz S. Alkalineearth metal compounds as alcoholysis catalysis for ester oils synthesis. Appl Catal A: General 2000;192:238. [61] Hattori H, Shima M, Kabashima H. Alcoholysis of ester and epoxide catalyzed by solid bases. Stud Surface Sci Catal 2000;130:350712. [62] DiSerio M, Tesser R, Ferrara A, Santacesaria E. Heterogeneous basic catalysts for the transesterication and the polycondensation reactions in PET production from DMT. J Mol Catal A: Chem 2004;212:2517. [63] Schuchardt U, Sercheli R, Vargas RM. Transesterication of vegetable oils: a review. J Braz Chem Soc 1998;9(1):199210. [64] Diasakov M, Loulodi A, Papayannakos N. Kinetics of the noncatalytic transesterication of soybean oil. Fuel 1998;77:1297302. [65] Barton AFM. Handbook of solubility and other cohesion parameters. Florida: CRC Press; 1985. [66] Saka S, Dadan K. Biodiesel fuel from rapeseed oil as prepared in supercritical methanol. Fuel 2001;80:22531. [67] Kusdiana D, Saka S. Kinetics of transesterication in rapeseed oil to biodiesel fuel as treated in supercritical methanol. Fuel 2001;80: 6938. [68] Al-Widyan MI, Al-Shyoukh AO. Experimental evaluation of the transesterication of waste palm oil into biodiesel. Bioresour Technol 2002;85:2536. [69] Kusdiana D, Saka S. Eects of water on biodiesel fuel production by supercritical methanol treatment. Bioresour Technol 2004;91: 28995. [70] Romano S. Vegetable oils-a new alternative. In: Vegetable oils fuels proceedings of the international conference on plant and vegetable oils as fuels. ASAE Publication 4-82. Fargo, ND, USA, 1982. p. 10116. [71] Demirbas A. Recent developments in biodiesel fuels. Int J Green Energy 2007;4:1526. [72] Naik M, Meher LC, Naik SN, Das LM. Production of biodiesel from high free fatty acid Karanja (Pongamia pinnata) oil. Biomass Bioenergy 2007. doi:10.1016/j.biombioe.2007.10.006. [73] Pasqualino JC, Montane D, Salvado J. Synergic eects of biodiesel in the biodegradability of fossil-derived fuels. Biomass Bioenergy 2006;30:8749. [74] Zhang X, Peterson C, Reece D, Moller G, Haws R. Biodegradability of biodiesel in the aquatic environment. Trans Am Soc Agric Eng 1998;41:142330. [75] Makareviciene V, Janulis P. Environmental eect of rapeseed oil ethyl ester. Renew Energy 2003;28(15):2395403. [76] Pereira G, Mudge SM. Cleaning oiled shores: laboratory experiments testing the potential use of vegetable oil biodiesels. Chemosphere 2004;54(3):297304. guez-Va zquez R, Herna ndez-Velasco M, [77] Molina-Barehona L, Rodr n C, Zapata-Pe rez O, Mendoza-Cantu A, et al. Appl Vega-Jarqu Soil Ecol 2004;27(2):16575. [78] Boopathy R. Anaerobic biodegradation of no. 2 diesel fuel in soil: a soil column study. Bioresour Technol 2004;94:14351. [79] Mohn WW, Radziminski CZ, Fortin MC, Reimer KJ. On site bioremediation of hydrocarbon-contaminated arctic tundra soils in inoculated biopiles. Appl Microbiol Biotechnol 2001;57:2427. [80] Mukherji S, Jagadevan S, Mohopatra G, Vijay A. Biodegradation of diesel oil by an arabian sea sediment culture isolated from the vicinity of an oil eld. Bioresour Technol 2004;95(3):2816.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

ARTICLE IN PRESS
Y.C. Sharma et al. / Fuel xxx (2008) xxxxxx [105] Demirbas A, Karslioglu S. Biodiesel production facilities from vegetable oils and animal fats. Energy Sources, Part A: Recovery, Utilization, Environ Eects 2007;29(2):13341. [106] www.re-focus.net. [107] Azam MM, Waris A, Nahar NM. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass Bioenergy 2005;29:293302. [108] Demirbas A. Importance of biodiesel as transportation fuel. Energy Policy 2007;35:466170. [109] Bender M. Economic feasibility review for community-scale farmer cooperatives for biodiesel. Bioresour Technol 1999;70:817. 19

[110] Canakci M, Van Gerpen J. A pilot plant to produce biodiesel from high free fatty acid feedstocks. Paper No. 016049, 2001 ASAE Annual International Meeting, Sacramento, CA. [111] Haas MJ, McAloon AJ, Yee WC, Foglia TA. A process model to estimate biodiesel production costs. Bioresour Technol 2006;97:6718. [112] Nas B, Berktay A. Energy potential of biodiesel generated from waste cooking oil: an environmental approach. Energy Sources, Part B: Econ Plan Policy 2007;2:6371. [113] Business Line, Business Daily from The Hindu group of publications, December 5, 2007.

Please cite this article in press as: Sharma YC et al., Advancements in development and characterization ..., Fuel (2008), doi:10.1016/ j.fuel.2008.01.014

Vous aimerez peut-être aussi