You are on page 1of 33

Graphical Abstract

A Review on Recent Biodiesel Production Methods Using Non-edible Oils

Chandra Has, Second Author, Third Author

Temperature controller

Alcohol Biodiesel
Power controller

Oil
Centrifuge/settling

Catalyst
Bi-products

Magnetic stirrer Time controller


Microwave cavity

Microwave reactor
Highlights
A Review on Recent Biodiesel Production Methods Using Non-edible Oils

Chandra Has, Second Author, Third Author

• Conventional methods of biodiesel production have been reviewed briefly.

• Various navel methods for the biodiesel production from non-edible oils.

• The role of the reaction conditions as well as the transport limitations in the
reaction.

• The influence of several parameters are reviewed.


A Review on Recent Biodiesel Production Methods
Using Non-edible Oils
Chandra Hasa,∗, Second Authora , Third Authorb
a
Department of Chemical Engineering, Indian Institute of Technology Bombay, Mumbai-400076,
India
b
Chemical Engineering group, CSIR, North East Institute of Science and Technology,
Jorhat-785006, India

Abstract

Among the various derivatives of vegetable oils and fats, fatty acid alkyl esters,
referred to as biodiesel fuel derived from triglycerides by the transesterification
reaction with alcohol, exhibit the promising alternative of diesel fuels and have re-
ceived the great attention nowadays. The physicochemical properties of biodiesel
have made the pursuit of high grade biodiesel production attractive. It has been
a great attention in almost all countries because it is nontoxic, biodegradable and
renewable diesel fuel. However, production cost of biodiesel on the commer-
cial purpose is a major hurdle. Using of non-edible oils or waste cooking oils
can be one of the good sources for the biodiesel production since such oils are
not used to for the purpose of human food consumption. This review not only
describes the conventional and novel production methods of biodiesel from non-
edible oils (mainly jatropha curcas oil) but also spotlights the role of the reaction
conditions as well as the transport limitations in the reaction system. Meanwhile,


Corresponding author
Email addresses: chandrahashbti@gmail.com (Chandra Has ),
author2email@gmail.com (Second Author), author3email@gmail.com (Third
Author)

Preprint submitted to Fuel July 22, 2020


the recent techniques, such as supercritical fluid, microwave irradiation, ultra-
sonic, microwave-ultrasonic, and co-solvent methods, significantly influence the
final conversion, yield and particularly, the product quality.
Keywords: Transesterification, esterification, microwave, ultrasonication,
supercritical, saponification

1 1. Introduction

2 For the human survival, energy consumption is inevitable. Today, most of


3 the energy consumption is achieved from fossil fuels, i.e., petroleum, coal, and
4 natural gas. Albeit fossil fuels are still being created by underground heat and
5 pressure, but they are being consumed more speedily than they are being created.
6 According to the statistical information provided by international agencies, the de-
7 mand of energy consumption will be increased by 53% by the year 2030. In USA
8 alone, it is predicted that the fossil fuel demand, especially petroleum products,
9 will increase from 84.40 to 116.00 million barrels/day by the year 2030 [1]. In an
10 analysis [2], it is reported that the fossil fuel reserve depletion times for oil, coal
11 and gas of approximately 35, 107, and 37 years, respectively. Due to these rea-
12 sons, fossil fuels are considered as non-renewable resources (once finished they
13 can not be created), and they will be depleted in the near future. Therefore, there
14 is a need to search an alternative fuel which should be technically feasible, eco-
15 nomically competitive, environmentally acceptable, and readily available. Various
16 alternative renewable energy sources such as biodiesel, biofuel, hydro, wind, so-
17 lar, nuclear, etc. have been suggested but unfortunately such sources are still in
18 the research and development stage [3]. Many researchers are concentrating on
19 the viability of vegetable oils (edible and non-edible oils), which are renewable,

2
20 biodegradable, easily available, and nontoxic. Moreover, their flash point (FP)
21 is relatively higher which makes them easy to store, transport, and handle. Var-
22 ious vegetable oils, waste cooking oils, and animal fats have been tested for this
23 purpose.
24 In the 19th century, the renowned scientist Sir Rudolf Christian Karl Diesel
25 (1858–1913), the inventor of diesel engine, demonstrated the use of edible/non-
26 edible oils as a substitute for diesel fuel in an exhibition in Paris [4]. With peanut
27 oil, an engine was tested for several hours. He was highly enthused, and asserted
28 that the time would come when the biomass fuel would be equally important
29 fuel. In the period of 1930s and 1940s, especially in the 2nd world war, biofu-
30 els were utilized in emergency to substitute diesel [5, 6]. In general, biodiesel
31 (Greek word, bio means life + diesel from Rudolf Christian Karl Diesel) refers to
32 a diesel-equivalent, processed fuel that is derived from various biological sources.
33 It can be defined as the monoalkyl esters of long-chain fatty acids which are de-
34 rived from renewable biolipids [7] such as edible [8], non–edible oils [9, 10],
35 or animal fats [11–13]. Albeit biodiesel has a ample of benefits in related to
36 petroleum diesel, but the expensiveness for its production is the main barrier to
37 its commercial use. In a study [14], price of biodiesel has been reported about
38 0.5 US$/L, compared to 0.35 US$/L for petroleum diesel. In particular, biodiesel
39 price depends on the cost of feedstocks, and this cost is almost 70–95% of the
40 total biodiesel cost [15–18]. The use of cheap edible, non-edible, or waste cook-
41 ing oils can be a way to enhance the economy of the biodiesel production and
42 its commercialization at the industry level. According to the climate conditions,
43 different countries have been adopting for different types of oils for the possible
44 use in the biodiesel production.

3
45 In the last few years, numerous review papers on the biodiesel production
46 from vegetable oils have been published, and many of them indicate the impor-
47 tance of non-edible oils in the biodiesel production. These articles have concen-
48 trated on many relevant topics for the utilization of non-edible oils in production
49 of the biodiesel. For example, some researchers have considered different tech-
50 niques to resolve the issues related with the high fuel viscosity [19–22], some
51 have described the physicochemical properties of non-edible vegetable oils and
52 esters produced from them [17, 19, 23, 24], and the engine performances with
53 both non-edible vegetable oils and biodiesels produced from non-edible sources
54 have been broadly reviewed by several authors [16, 21, 25–27]. Moreover, in
55 order to reduce free fatty acids (FFAs) in various non-edible oils, some papers
56 provide an insightful information to use catalytic and non-catalytic transesterifi-
57 cation processes [1, 16, 18, 19, 21, 27–30]. Apart from these mentioned topics,
58 the impact of biodiesel on the environment, and the production cost of biodiesel
59 have been touched in a few reviews [24, 31]. Many studies have concentrated on
60 the biodiesel production from jatropha oil in India [32], Brazil [33], China [34],
61 Malaysia [35], and Indonesia [36]. In a review [37], only ten countries (Malaysia,
62 Indonesia, Argentina, USA, Brazil, Germany, Philippines, Belgium, Spain) pro-
63 duce more than 80% of the total biodiesel production.
64 Present article reviews all possible conventional and novel methods in the
65 production of biodiesel from vegetable oils, particularly from jatropha oil. Al-
66 though several articles are already published based on the conventional and novel
67 biodiesel production methods, but the area of the biodiesel research has become
68 very vast in the last some years, therefore several methods have been developed
69 with the time and some are in the pipeline. It has been tried to cover more and

4
70 more novel methods in this review. In addition, other major goal of this article
71 is to present the possibilities of the use of non-edible oils (e.g., jatropha oil) in
72 the biodiesel production, to consider the major approach for the treatment of non-
73 edible oils before transesterification process, and to study the influence of some
74 operating parameters and reaction conditions on the production rate.

75 2. Conventional biodiesel production method

76 There are various kinds of modification processes, e.g., dilution or blending,


77 thermal cracking or pyrolysis, transesterification, and microemulsification, and
78 among which transesterification (also known as alcoholysis) is found to be the
79 best conventional method for producing higher quality of biodiesel [3]. We will
80 briefly describe each conventional method.

81 2.1. Dilution method

82 In dilution method, vegetable oils are diluted with petroleum diesel to run the
83 engine. In this way, high viscosity of vegetable oils are minimized by blending
84 the oils with petroleum diesel; method does not need any chemical process [38].
85 Caterpillar Brazil, in the year 1980, used a mixture of vegetable oil and petro-
86 diesel in the ratio of 1:9, and this mixture was tested in pre-combustion chamber
87 engine without any modification or adjustment to the engine. At that time, it was
88 not practically possible to use 100% vegetable oil for diesel fuel, however, a blend
89 of 20% vegetable oil and 80% diesel fuel was successfully working for short-term
90 use [11].

5
91 2.2. Pyrolysis

92 Pyrolysis refers to the conversion of one species into another by the application
93 of heat or by heat with the aid of a catalyst in the absence of air or oxygen [39].
94 The pyrolyzed substance can be vegetable edible or non-edible oils, animal fats,
95 fatty acids, and fatty methyl esters. [40] studied the catalytic cracking of vegetable
96 oils to produce biofuels. In a review by [11], cracking of copra and palm oils at
97 450 ◦ C has been studied in which gases, liquids and solids with lower molecular
98 weights have been observed, and the condensed organic phase is fractionated to
99 produce biogasoline and biodiesel fuels. This method is not only expensive but
100 also it produces some low value substances and, sometimes, more gasoline than
101 diesel fuel.

102 2.3. Microemulsifications

103 In this method, high viscosity of vegetable oils is reduced by making mi-
104 croemulsions with solvents such as methanol, ethanol, and n-butanol. Microemul-
105 sions are the stable colloidal dispersions in which size of the dispersed-phase
106 particles is less than one-fourth the wavelength of visible light. Fuels based on
107 microemulsions are sometimes also referred to as “hybrid fuels”, albeit blends of
108 petro-diesel with vegetable oils are also termed as hybrid fuels [41].

109 2.4. Transesterification process

110 Transesterification process is a conventional and the most notable method for
111 the biodiesel production. In this reaction, vegetable oils react with short chain pri-
112 mary alcohol and produce fatty alkyl esters (biodiesel) and glycerol (bi-product).
113 Normally, an electrical heater (or heating coils) and an electrical motor are em-
114 ployed for the purpose of heating and mixing of the reactants, respectively. The

6
115 heating merely depends on the convection currents, and the thermal conductivity
116 of the reaction mixture. The reaction temperature is, generally, maintained below
117 the boiling point of alcohol.
118 It can be performed with or without any types of catalyst (alkali, acid, and
119 enzymatic catalyst) [42]. In general, catalysts split the feedstock into glycerin and
120 biodiesel and they can make production easier and faster. In a study [11], time
121 consumed in tansesterification reaction with alkali catalyst is observed shorter
122 than any other types of catalyzed reaction. It has been reported more than 97%
123 conversion yield in the case of alkali catalyzed transesterification reaction [43].
124 However, for higher yield of conversion, it is very important that the reactants (oil
125 and alcohol) should be significantly anhydrous [44] otherwise water will reduce
126 the formation of fatty acid alkyl ester (FAAE) during the transesterification reac-
127 tion [45]. As discussed, alkali catalyst has been found to be best choice over other
128 catalysts but it has even several drawbacks in comparison to lipase catalysts, as
129 shown in Table 1.
130 In this process, fatty esters are formed by the reaction of triglycerides with an
131 alcohol, especially short chain alcohols such as ethanol or methanol, in the pres-
132 ence of alkali, acid or enzyme catalyst etc. First, catalyst is dissolved in alcohol
133 (taking example of methanol) and then the mixture is contacted to the feedstock
134 (oil/fat) in the stirred reactor. For achieving reaction equilibrium rapidly, reac-
135 tor temperature is maintained around 55 ◦C to 60 ◦C. 3 moles of methanol react
136 with one mole of triglyceride and produce 3 moles of methyl esters with one
137 mole of glycerol as a by-product (Figure 1). Since reaction is reversible, there-
138 fore for shifting equilibrium in the forward direction, methanol is used in the
139 excess amount. For alkali catalyzed transesterification reaction, 6:1 molar ratio

7
Table 1: Comparative study of alkali and enzymatic catalyzed process for the biodiesel production
(adapted from [46]).

State Alkali catalyzed Lipase catalyzed

Operating Temp. 60 ◦C to 70 ◦C 30 ◦C to 40 ◦C
FFA in crude oil Saponified product Methyl esters
Yield of methyl esters Normal Higher
Recovery of glycerol Difficult Easy
Water in raw material Interference with reaction No influence
Catalyst requirement Low High
Purification of methyl esters Repeated washing None
Production cost of catalyst Cheap Relatively expensive

140 of methanol to oil gives the sufficient conversion of oil to biodiesel [47, 48]. The
141 preferable quantity of alkali catalyst for transesterification reaction can be taken
142 between 0.1 to 1% (w/w) of oils and fats [8, 11]. In the case of ethanol, fatty acid
143 ethyl ester (biodiesel) and glycrol are produced as products. However, methyl
144 ester is preferred over ethyl ester because methyl ester is the predominant prod-
145 uct of commerce, highly reactive, and other advantage is that it is considerably
146 cheaper compared to ethanol. In addition to alcohols, propanol, butanol, and amyl
147 alcohol are the other alcohols which can be employed for the transesterification
148 process [49].

149 2.4.1. Factors affecting transesterification reaction


150 There are several factors which can affect the transesterification reaction such
151 as reaction temperature, high FFA, moisture content, molar ratio of alcohol to oil,
152 catalyst type, and mass transfer rate.

8
O O

CH2 O C R1 CH3 O C R1 CH2 OH


O O
Catalyst
CH O C R2 + 3 CH3 OH CH3 O C R2 + CH OH
O O

CH2 O C R3 CH3 O C R3 CH2 OH


Triglyceride Methanol Methyl Esters (Biodiesel) Glycerol

Figure 1: Transesterification of triglyceride with methanol in the presence of alkali catalyst (R1 ,
R2 and R3 are the different alkyl groups)

O O
NaOH
CH3 O C R + H2 O Catalyst R C OH + CH3 OH
(Methyl ester) (Water) (FFA) (Methanol)
Figure 2: Hydrolysis of ester due to presence of water (a side reaction which produces free fatty
acid)

153 Effect of water. It is always suggested that all reactants for the tranesterification
154 should be anhydrous because water has a negative impact on the reaction. Water
155 can inhibit the catalyst efficiency by absorbing the catalyst. It has been that the
156 presence of water in feedstock can have more negative impact on reaction than
157 that of free fatty acids [50]. Water enhances the soap formation by a side reaction
158 so the resulting soap increases the viscosity and hence separation of glycerol from
159 products becomes more difficult. A side reaction due to water has been shown
160 in Figure 2. Water produces FFA and eventually FFA reacts with alkali catalyst
161 which forms more soap and water unless catalyst is consumed and deactivated.

162 Effect of reaction time. In general, conversion rate of tryglycerides into esters
163 increases with increasing reaction time [11, 30]. Freedman and co-workers inves-
164 tigated the effect of reaction time under the condition of methanol to oil molar

9
165 ratio of 6:1, 0.5% NaOCH3 catalyst, and 60 ◦ C temperature. At the beginning,
166 due to the mixing and dispersion of alcohol into the feedstock the reaction is slow,
167 but after a while the reaction proceeds very rapidly [51]. Usually, the conversion
168 attains a maximum at a certain reaction time (¡ 90 min), and after that conver-
169 sion remains almost invariant with a further increase in the reaction time [52, 53].
170 In addition to the product yield, excess reaction time leads to a decrease in the
171 biodiesel formation due to the backward transesterification reaction, resulting not
172 only in a loss of the yield (fatty alkyl esters) but also causing more fatty acids to
173 the formation of soaps [54].

174 Effect of free fatty acid. It is firmly asserted that the FFA in oil should not be
175 more than 1% (which corresponds acid value of 2 mg KOH/g oil) for base cat-
176 alyzed transesterification reaction [29, 42, 47, 55]. FFA above this level causes
177 soap formation with alkali catalyst which obstacles the separation of ester from
178 the glycerol and rate of ester conversion is also reduced. Moreover, the catalyst
179 activity is inhibited due to soap formation, therefore, the yield of biodiesel is re-
180 duced.

181 Effect of reaction temperature. Rate of transesterification reaction is significantly


182 influenced by reaction temperature. However, transesterification of oil with alkali
183 catalyst gives the satisfactory biodiesel yield if reaction proceeds for a long time at
184 the room temperature. Usually temperature is maintained between 60 ◦C to 70 ◦C
185 (close to boiling point of methanol). Temperature above this will burn methanol
186 and therefore yield of biodiesel will reduce. At higher temperature, researches
187 noted negative impact on the yield of transesterification of neat oil because high
188 temperature increases the side saponification reaction rate of triglyceride. Later
189 on, optimum yield was found at 60 ◦C in just reaction time of 20 min [52].

10
190 Effect of molar ratio. It is one of the most important factor which influences the
191 yield of esters during the transesterification reaction. The stoichiometric molar
192 ratio of alcohol to triglyceride is 3:1. Molar ratio closely depends on the type of
193 catalyst used. Transesterification reaction is a reversible reaction so for shifting
194 the reaction to direction of ester formation, alcohol is used in excess amount or
195 one of the product is removed from the reaction mixture. For alkali catalyzed
196 transesterification reaction, most common molar ratio of pretreated oil (FFA ¡
197 1%) to methanol is 6:1 which is used in industrial purpose in order to obtain more
198 than 98% methyl ester on a weight basis.

199 Effect of catalyst type and concentration. Alkali, acid, and enzyme catalysts are
200 very common for the biodiesel production. In one study, alkali catalyzed transes-
201 terification was much faster and less corrosive than acid catalyzed transesterifica-
202 tion. Acid catalyzed transesterification is more preferable if oil contains high FFA
203 but in this case reaction will require higher temperature and longer reaction time
204 than base catalyzed transesterification. Sodium methoxide is more effective than
205 sodium hydroxide because NaOH produces water with alcohol [51]. However,
206 sodium and potassium hydroxide are also used as a base catalyst for the trans-
207 esterification, and due to their low cost, they are primarily chosen in industrial
208 biodiesel processing [46]. Lipase catalyzed transesterification can also be a good
209 choice but the production cost is significantly greater than alkaline one. Catalyst
210 concentration in the range of 0.5–1% (by weight of oil) is the most suitable choice
211 for the biodiesel processing [8]. Some authors suggest to use any value between
212 0.005–0.35% (w/w) [11].

11
213 3. Recent biodiesel production methods

214 3.1. Supercritical production method for biodiesel

215 Biodiesel can be easily produced by the supercritical process without using of
216 any catalyst. A supercritical fluid is referred to as any substance at a temperature
217 and pressure above its critical point. These fluids are environment amenable and
218 economical. Figure 3 illustrates the general biodiesel production by continuous su-
219 percritical alcohol process. The yield of biodiesel with the supercritical methanol
220 method is found to be higher than those obtained in the conventional method with
221 a basic catalyst; conventional and supercritical methanol methods are compared
222 in Figure 3. Methanol in liquid phase is a polar solvent and it has hydrogen bond-
223 ing between OH-oxygen and OH-hydrogen to forms methanol clusters, whereas
224 supercritical methanol has a hydrophobic nature with a lower dielectric constant,
225 therefore triglyceride, which is non-polar in nature, can be well solvated in the
226 presence of supercritical methanol in order to form a single phase oil/methanol
227 mixture. Due to this reason, the conversion rate of oil to methyl ester is found to
228 increase dramatically in the supercritical state.
229 In general, supercritical methanol and supercritical ethanol are highly used
230 for the biodiesel production. Nevertheless, supercritical carbon dioxide (ScCO2 )
231 can be also employed for this purpose because it is a cheaper, inflammable, and
232 non-toxic solvent [56, 57]. In one recent study [58], two-step transesterification
233 methods, such as both subcritical and supercritical, and both enzyme and super-
234 critical fluid conditions etc. have been also developed. In spite of the same ben-
235 efits as one-step supercritical methanol method, the two-step process is found to
236 use milder reaction condition and shorter reaction time. Furthermore, the method
237 enables to use common stainless steel for the manufacturing of reactor and lower

12
Temperature indicator
Alcohol recovery

Oil Alcohol

Condenser

High pressure pump Oven

Pressure indicator Biodiesel

Figure 3: Production of biodiesel by continuous supercritical alcohol method.

238 the energy consumption [59]. The two-step supercritical method has several ad-
239 vantages, for example, no soap formation, no pollution, easy in the purification,
240 no catalyst, no waste water, no equipment high cost, no high alcohol/oil ratio.
241 This method can further conform to environmental concerns if volatile, toxic, and
242 flammable solvents are avoided and replaced the enzyme with ScCO2 [60].

243 3.2. Microwave assisted technique

244 Usually, heating coils are employed to heat the raw material in the biodiesel
245 production method. In several studies, it is stated that the microwave heating can
246 also be employed for the transesterification. The microwave frequencies can be
247 varied from 300 MHz to 30 GHz, however, frequency of 2.45 GHz is preferred
248 in the laboratory applications [61]. The main advantage of microwave heating
249 method is that a very high temperatures can be achieved in a short periods of
250 time [62].
251 In this method, the reactants, such as vegetable oil, alcohol, and catalyst, are

13
252 well mixed using suitable agitator and heated by the microwave heat source; trans-
253 esterification reaction takes place. The reactants are operated under the microwave
254 irradiation in order to achieve the reaction time of few min [63]. Nüchter and
255 co-workers describe the application of microwave treatment for selected organic
256 reactions. In their study, microwaves activate the smallest degree of variance of
257 polar molecules and ions which lead to molecular friction, and hence, the initia-
258 tion of chemical reactions is possible [64]. Since the interaction of energy with the
259 sample occur at a molecular level, very efficient and fast heating can be achieved
260 by means of microwave heating. The molecules do not have a sufficient time to
261 relax due to very quick energy interaction with sample molecules. In this way,
262 the generated heat can be for short times and much greater than the overall set
263 temperature of the bulk reaction mixture [65, 66]. Transesterified products are
264 then allowed to settle for the phase separation. The two phases, i.e., biodiesel and
265 glycerin, are separated within 30–60 min. After the separation, crude biodiesel
266 is treated with water wash to separate the impurities and dried to remove the
267 moisture. The reaction and settling times in this method are low and therefore
268 the production cost also decreases significantly [67]. Figure 4 shows the simple
269 schematic of microwave technique of the biodiesel production.

270 3.3. Ultrasound assisted technique

271 Ultrasonic or ultrasound waves are the energy application of sound waves
272 which is vibrated more than 20 kHz or 20000 cycles per second [68]. Accord-
273 ing to the some investigators, ultrasonic irradiation has three effects: first, rapid
274 migration of fluid particles due to a variation of sonic pressure, second is micro-
275 scopic bubbles (cavities), and third one is acoustic streaming mixing. Cavities are
276 created when liquid breaks down under the application of large negative pressure

14
Temperature controller

Alcohol Biodiesel
Power controller

Oil
Centrifuge/settling

Catalyst
Bi-products

Magnetic stirrer Time controller


Microwave cavity

Microwave reactor

Figure 4: Schematic representation of microwave assisted transesterification technique.

277 gradient. When ultrasonic intensities (watts/cm2 ) are high, small cavity may grow
278 quickly due to inertial effects, therefore, bubbles grow and implode violently. The
279 generation and collapse of small bubbles are responsible for most of the chemical
280 effects. Note that cavitation enhances the rate of mass transfer by disrupting the
281 interfacial boundary layers, this is referred to as “jet effct” [69]. Due to a high en-
282 ergy action, the reactivity of the reactants remarkably enhances and reaction time
283 greatly reduces without involving elevated temperatures. In fact, such reaction
284 can be processed at or slightly above the ambient temperature. The highest yield
285 of biodiesel has been observed at an ultrasonic power above of 450 W [70].
286 The ultrasonic irradiation is a very desirable technique to produce biodiesel
287 from vegetable oils and animal fats, because it reduces the processing cost, in-
288 creases the transesterification reaction rate, does not demand high temperature,
289 and produces a high grade biodiesel. Moreover, it also facilities the change of pro-
290 duction process from batch method to continuous flow processing. Furthermore,
291 current users of this method claim that much less amount of catalyst and methanol
292 are needed. This process, in contrast, has few disadvantages, e.g., slightly higher

15
Ultrasonic probe Ultrasonic generator

Oil
Biodiesel

Phase separator
Alcohol
Alcohol

Mixer M
ix Reactor Separator
er
Catalyst Glycerol

Figure 5: Schematic diagram of ultrasonic-assisted batch reactor system for the biodiesel pro-
duction. The ultrasonic probe enables the reactor with ultrasonic energy followed by the phase
separation stage. In the phase separation stage, biodiesel can be collected whereas the remaining
residue can be further subjected to extract the catalyst and glycerin (adapted from [72]).

293 reaction temperature for long reactions, higher catalyst loading in comparison to
294 conventional method, and the ultrasonic power should be under control due to the
295 soap formation in rapid reaction. Biodiesel yield, therefore, may be decreased by
296 the application of higher frequencies (40 kHz).
297 In this process, first, vegetable oil is mixed with alcohol and catalyst, and then
298 the mixture is heated. The heating mixture is then sonicated inline, and in the last,
299 glycerol is separated using a centrifuge (see Figure 5). Some researchers have
300 used alternative reactors to reduce the power consumption. For example, Cintas
301 and co-workers developed a flow reactor constituted with three transducers. The
302 authors demostrated considerable amount of energy saving by large-scale multi-
303 ple transducer sonochemical reactors employing in a continuous mode [71]. The
304 factors, such as catalyst and alcohol types, ultrasound power, and frequency, can
305 influence the rate of the biodiesel production in this method.

16
306 3.4. Microwave-ultrasound assisted technique

307 In some studies, biodiesel is produced using combined techniques of microwave


308 and ultrasound. At an optimum conditions, the reaction time for the esterification
309 and transesterification using ultrasound alone is reported as 60 min and 20 min,
310 respectively and this reaction time is reduced to only 15 min and 6 min for the se-
311 quential approach. The optimal steps include as ultrasonic mixing for 1-min and
312 closed microwave irradiation for 2-min, and the optimal reaction conditions are set
313 for amount of catalyst (1.0 wt%), operating temperature (30 ◦ C), and methanol-oil
314 (6:1, by mol) [73]. Yin and co-workers have compared four different enhancing
315 methods, i.e., mechanical stirring (MS), flat plate ultrasonic irradiation (FPUI),
316 flat plate ultrasonic irradiation with mechanical stirring (UIMS), and probe ultra-
317 sonic irradiation (PUI), in order to select a better one that need minimal amount
318 of catalyst, less energy consumption and time to attain equilibrium for producing
319 biodiesel through transesterification process of sunflower oil [74]. Their results
320 describe that under the same conditions, UIMS and PUI require less amount of
321 catalyst, less alcohol, shorter reaction time and less energy consumption in com-
322 parision to the MS and FPUI with the same biodiesel conversion.

323 3.5. Co-solvent assisted technique

324 Another method applied in the biodiesel production is a co-solvent assisted


325 method (using of THF or other co-solvents), which enhances the feedstock mis-
326 cibility with alcohol [75–77]. In this method, 1-phase oil rich system is formed,
327 and with the use of inert co-solvents the triglycerides are converted into esters.

17
328 4. Conclusions

329 In the last few years, biodiesel has been a great attention as an alternative
330 fuel for diesel engines because of its biodegradable benefits, non-toxicity, and the
331 fact that it is produced from renewable resources. Cheap edible oils, non-edible
332 oils, waste cooking oils, and animal fats can be good sources used to supplement
333 mineral diesel oil. Fortunately, engines are required no modifications to replace
334 the diesel fuel with such biofuels. Biodiesel is produced by the most notable
335 method which is known as transesterification process. Different approaches for
336 the biodiesel production, such as acid, alkali, or enzymatic catalyzed, and non-
337 catalyzed transesterification processes have been tried. All of them, however,
338 have their own pros and cons. The type of feedstock is the most considerable
339 factor in the biodiesel production. Utilizing non-edible oils can be a good op-
340 tion for the biodiesel production since edible oils are consumed as human food.
341 Jatropha curcas, a non-edible source, has been a good source of the biodiesel pro-
342 duction because it is a drought-resistant perennial and can grow on barren land
343 and also it requires extremely low amount of water. The main purpose of the
344 transesterification is to lower the viscosity of oils or fats. Before alkali catalyzed
345 transesterification, high free fatty acid (FFA) is reduced to less than 1% by esteri-
346 fication reaction with acid catalyst. Study shows that a very small amount of water
347 in oil can produce more negative impact than high FFA. In addition, the biodiesel
348 production from the conventional alkali catalyzed transesterification process is
349 much longer. Therefore, various methods, such as supercritical, microwave, ul-
350 trasonic, microwave-ultrasonic, and co-solvent to decrease the production costs,
351 reaction time, catalyst and alcohol, have been employed in the transesterification
352 reactions. Such novel methods can enhance the quality of fatty methyl esters for

18
353 the applications to diesel engines without any kind of modification in the engine.

354 Appendix A. Sources of biodiesel

355 Edible oils are the major resources of the biodiesel production. Today, ∼
356 84% of the world biodiesel production is achieved from rapeseed oil. The re-
357 maining parts are achieved from sunflower oil (13%), palm oil (1%), soybean
358 oil, and from others (2%), respectively [17]. More than 95% of world biodiesel
359 is produced from edible oils because these are easily available from the agricul-
360 tural plants [17]. Usually, the biodiesel sources depends on the susceptibility of
361 crops to the regional climate conditions. For instance, in USA, soybean oil is
362 the most commonly biodiesel feedstock, and on the other hand, in Europe and
363 in tropical countries, rapeseed (or canola) and palm oils are the most common
364 feedstocks for the biodiesel production [78]. However, it is great concern to pro-
365 duce biodiesel from edible oils because they compete our food requirements. It
366 may bring global imbalance to the demand and market supply due to high edible
367 oil price, reduction of sources and growth of population. These issues, therefore,
368 can be resolved by using of cheaper and non-edible vegetable oils since they do
369 not compete with edible vegetable oils on the food market, or waste cooking can
370 also be used. For instance, mahua (Madhuca indica) [79], jatropha tree (Jatropha
371 curcas) [80], karanja (Pongamia pinnata) [81], algae [82, 83], microalgae [84],
372 and grease oil [85] can be employed. The major hurdle in the industrialization
373 and commercialization of biodiesel is the production costs. To concern this issue,
374 waste edible oils can be used since they can reduce biodiesel production costs by
375 60–90% [14, 86, 87].
376 Among the aforementioned non-edible sources, jatropha oil is one of the major

19
Table A.2: Oil content and production of non edible oil seeds (adapted from [88]).

Species Oil fraction Seed yield Oil yield


(%) (×106 tones/year) (tones/year)

Jatropha 50–60 0.20 2.0–3.0


Mahua 35–40 0.20 1.0–4.0
Karanja 30–40 0.06 2.0–4.0
Castor 45–50 0.25 0.5–1.0
Linseed 35–45 0.15 0.5–1.0
Others 10–50 0.50 0.5–2.0

377 non-edible source in many countries; major non-edible oils are compared in Ta-
378 ble A.2. It can be extracted from jatropha curcas which is a drought-resistant
379 perennial and growing well in marginal and barren land; it is much less expen-
380 sive than edible sources and that is why it is being considered as a promising
381 alternaitve to edible vegetable oils. Seeds of the jatropha curcas are the primary
382 source for extraction of oil.

383 Acknowledgments

384 The authors would like to thank Mr. John Smith for his expertise and assis-
385 tance throughout all aspects of our study and for his help in writing the manuscript.

386 References

387 [1] E. M. Shahid, Y. Jamal, Production of biodiesel: a technical review, Renew.


388 Sust. Energ. Rev. 15 (9) (2011) 4732–4745.

20
389 [2] S. Shafiee, E. Topal, When will fossil fuel reserves be diminished?, Energy
390 Policy 37 (1) (2009) 181–189.

391 [3] A. Talebian-Kiakalaieh, N. A. S. Amin, H. Mazaheri, A review on novel


392 processes of biodiesel production from waste cooking oil, Appl. Energy 104
393 (2013) 683–710.

394 [4] B. Orchard, J. Denis, J. Cousins, Developments in biofuel processing tech-


395 nologies, World Pumps 2007 (487) (2007) 24–28.

396 [5] G. P. A. G. Pousa, A. L. F. Santos, P. A. Z. Suarez, History and policy of


397 biodiesel in Brazil, Energy Policy 35 (11) (2007) 5393–5398.

398 [6] H. Kim, B. Kang, M. Kim, Y. M. Park, D. Kim, J. Lee, K. Lee, Transesterifi-
399 cation of vegetable oil to biodiesel using heterogeneous base catalyst, Catal.
400 Today 93 (2004) 315–320.

401 [7] D. Ayhan, Biodiesel a Realistic Fuel Alternative for Diesel Engines, 1st edi-
402 tion Springer-Verlag London, 2008.

403 [8] B. K. Barnwal, M. P. Sharma, Prospects of biodiesel production from veg-


404 etable oils in India, Renew. Sust. Energ. Rev. 9 (4) (2005) 363–378.

405 [9] I. B. Banković-Ilić, O. S. Stamenković, V. B. Veljković, Biodiesel production


406 from non-edible plant oils, Renew. Sust. Energ. Rev. 16 (6) (2012) 3621–
407 3647.

408 [10] A. Murugesan, C. Umarani, T. R. Chinnusamy, M. Krishnan, R. Subrama-


409 nian, N. Neduzchezhain, Production and analysis of bio-diesel from non-
410 edible oils: a review, Renew. Sust. Energ. Rev. 13 (4) (2009) 825–834.

21
411 [11] F. Ma, M. A. Hanna, Biodiesel production: a review, Bioresour. Technol.
412 70 (1) (1999) 1–15.

413 [12] M. Gürü, B. D. Artukoğlu, A. Keskin, A. Koca, Biodiesel production from


414 waste animal fat and improvement of its characteristics by synthesized nickel
415 and magnesium additive, Energy Convers. Manag. 50 (3) (2009) 498–502.

416 [13] L. Canoira, M. Rodrı́guez-Gamero, E. Querol, R. Alcántara, M. Lapuerta,


417 F. Oliva, Biodiesel from low-grade animal fat: production process assess-
418 ment and biodiesel properties characterization, Ind. Eng. Chem. Res. 47 (21)
419 (2008) 7997–8004.

420 [14] Y. Zhang, M. A. Dube, D. D. l. McLean, M. Kates, Biodiesel production


421 from waste cooking oil: 1. Process design and technological assessment,
422 Bioresour. Technol. 89 (1) (2003) 1–16.

423 [15] M. Balat, Potential alternatives to edible oils for biodiesel production: a
424 review of current work, Energy Convers. Manag. 52 (2) (2011) 1479–1492.

425 [16] X. Fan, R. Burton, Recent development of biodiesel feedstocks and the ap-
426 plications of glycerol: a review, Open Fuel Energ. Sci. J. 2 (2009) 100–109.

427 [17] M. M. Gui, K. T. Lee, S. Bhatia, Feasibility of edible oil vs. non-edible oil vs.
428 waste edible oil as biodiesel feedstock, Energy 33 (11) (2008) 1646–1653.

429 [18] D. Y. C. Leung, X. Wu, M. K. H. Leung, A review on biodiesel production


430 using catalyzed transesterification, Appl. Energy 87 (4) (2010) 1083–1095.

431 [19] A. Demirbas, Progress and recent trends in biodiesel fuels, Energy Convers.
432 Manag. 50 (1) (2009) 14–34.

22
433 [20] M. Balat, H. Balat, Progress in biodiesel processing, Appl. Energy 87 (6)
434 (2010) 1815–1835.

435 [21] I. Kralova, J. Sjöblom, Biofuels–renewable energy sources: a review, J. Dis-


436 persion Sci. Technol. 31 (3) (2010) 409–425.

437 [22] N. N. A. N. Yusuf, S. K. Kamarudin, Z. Yaakub, Overview on the current


438 trends in biodiesel production, Energy Convers. Manag. 52 (7) (2011) 2741–
439 2751.

440 [23] A. Karmakar, S. Karmakar, S. Mukherjee, Properties of various plants and


441 animals feedstocks for biodiesel production, Bioresour. Technol. 101 (19)
442 (2010) 7201–7210.

443 [24] A. Kumar, S. Sharma, Potential non-edible oil resources as biodiesel feed-
444 stock: an Indian perspective, Renew. Sust. Energ. Rev. 15 (4) (2011) 1791–
445 1800.

446 [25] I. M. Atadashi, M. K. Aroua, A. A. Aziz, High quality biodiesel and its
447 diesel engine application: a review, Renew. Sust. Energ. Rev. 14 (7) (2010)
448 1999–2008.

449 [26] S. A. Basha, K. R. Gopal, S. Jebaraj, A review on biodiesel production,


450 combustion, emissions and performance, Renew. Sust. Energ. Rev. 13 (6)
451 (2009) 1628–1634.

452 [27] S. Y. No, Inedible vegetable oils and their derivatives for alternative diesel
453 fuels in CI engines: a review, Renew. Sust. Energ. Rev. 15 (1) (2011) 131–
454 149.

23
455 [28] Z. Helwani, M. R. Othman, N. Aziz, W. J. N. Fernando, J. Kim, Technolo-
456 gies for production of biodiesel focusing on green catalytic techniques: a
457 review, Fuel Process. Technol. 90 (12) (2009) 1502–1514.

458 [29] M. Y. Koh, T. I. M. Ghazi, A review of biodiesel production from jatropha


459 curcas L. oil, Renew. Sust. Energ. Rev. 15 (5) (2011) 2240–2251.

460 [30] A. P. Vyas, J. L. Verma, N. Subrahmanyam, A review on FAME production


461 processes, Fuel 89 (1) (2010) 1–9.

462 [31] J. Janaun, N. Ellis, Perspectives on biodiesel as a sustainable fuel, Renew.


463 Sust. Energ. Rev. 14 (4) (2010) 1312–1320.

464 [32] S. Jain, M. P. Sharma, Prospects of biodiesel from Jatropha in India: a re-
465 view, Renew. Sust. Energ. Rev. 14 (2) (2010) 763–771.

466 [33] D. M. Lapola, J. A. Priess, A. Bondeau, Modeling the land requirements and
467 potential productivity of sugarcane and jatropha in Brazil and India using the
468 LPJmL dynamic global vegetation model, Biomass Bioenerg. 33 (8) (2009)
469 1087–1095.

470 [34] C. Y. Yang, Z. Fang, B. Li, Y. Long, Review and prospects of Jatropha
471 biodiesel industry in China, Renew. Sust. Energ. Rev. 16 (4) (2012) 2178–
472 2190.

473 [35] M. Mofijur, H. H. Masjuki, M. A. Kalam, M. A. Hazrat, A. M. Li-


474 aquat, M. Shahabuddin, M. Varman, Prospects of biodiesel from jatropha
475 in Malaysia, Renew. Sust. Energ. Rev. 16 (7) (2012) 5007–5020.

24
476 [36] A. S. Silitonga, A. E. Atabani, T. M. I. Mahlia, H. H. Masjuki, I. A. Badrud-
477 din, S. Mekhilef, A review on prospect of Jatropha curcas for biodiesel in
478 Indonesia, Renew. Sust. Energ. Rev. 15 (8) (2011) 3733–3756.

479 [37] M. Johnston, T. Holloway, A global comparison of national biodiesel pro-


480 duction potentials, Environ. Sci. Technol. 41 (23) (2007) 7967–7973.

481 [38] V. S. Hariharan, K. V. Reddy, K. Rajagopal, Study of the performance, emis-


482 sion and combustion characteristics of a diesel engine using sea lemon oil-
483 based fuels, Indian J. Sci. Technol. 2 (4) (2009) 43–47.

484 [39] N. O. V. Sonntag, Structure and composition of fats and oils, in: D. Swern
485 (Ed.), Bailey’s industrial oil and fat products, 4th Edition, Vol. 1, New York:
486 John Wiley and Sons, 1979, pp. 289–477.

487 [40] D. Pioch, P. Lozano, M. C. Rasoanantoandro, J. Graille, P. Geneste,


488 A. Guida, Biofuels from catalytic cracking of tropical vegetable oils,
489 Oleagineux 48 (6) (1993) 289–292.

490 [41] G. Knothe, R. O. Dunn, M. O. Bagby, Biodiesel: the use of vegetable oils
491 and their derivatives as alternative diesel fuels, in: ACS symposium series,
492 Vol. 666, Washington, DC: American Chemical Society, 1997, pp. 172–208.

493 [42] G. G. Kombe, A. K. Temu, H. M. Rajabu, G. D. Mrema, High free fatty


494 acid (FFA) feedstock pre-treatment method for biodiesel production, in: Sec-
495 ond International Conference on Advances in Engineering and Technology,
496 2012, pp. 176–182.

25
497 [43] A. K. Singh, S. D. Fernando, R. Hernandez, Base-catalyzed fast transester-
498 ification of soybean oil using ultrasonication, Energy Fuels 21 (2) (2007)
499 1161–1164.

500 [44] H. J. Wright, J. B. Segur, H. V. Clark, S. K. Coburn, E. E. Langdon, R. N.


501 DuPuis, A report on ester interchange, Oil Soap 21 (5) (1944) 145–148.

502 [45] A. Demirbas, Biodiesel production via non-catalytic SCF method and
503 biodiesel fuel characteristics, Energy Convers. Manag. 47 (15–16) (2006)
504 2271–2282.

505 [46] H. Fukuda, A. Kondo, H. Noda, Biodiesel fuel production by transesterifica-


506 tion of oils, J. Biosci. Bioeng. 92 (5) (2001) 405–416.

507 [47] A. K. Tiwari, A. Kumar, H. Raheman, Biodiesel production from jatropha oil
508 (Jatropha curcas) with high free fatty acids: an optimized process, Biomass
509 Bioenerg. 31 (8) (2007) 569–575.

510 [48] H. Lu, Y. Liu, H. Zhou, Y. Yang, M. Chen, B. Liang, Production of biodiesel
511 from Jatropha curcas L. oil, Comput. Chem. Eng. 33 (5) (2009) 1091–1096.

512 [49] S. L. Suib, New and future developments in catalysis: catalytic biomass
513 conversion, Newnes, 2013.

514 [50] D. Kusdiana, S. Saka, Effects of water on biodiesel fuel production by su-
515 percritical methanol treatment, Bioresour. Technol. 91 (3) (2004) 289–295.

516 [51] B. Freedman, E. H. Pryde, T. L. Mounts, Variables affecting the yields of


517 fatty esters from transesterified vegetable oils, J. Am. Oil Chem. Soc. 61 (10)
518 (1984) 1638–1643.

26
519 [52] D. Y. C. Leung, Y. Guo, Transesterification of neat and used frying oil: op-
520 timization for biodiesel production, Fuel Process. Technol. 87 (10) (2006)
521 883–890.

522 [53] O. J. Alamu, M. A. Waheed, S. O. Jekayinfa, T. A. Akintola, Optimal trans-


523 esterification duration for biodiesel production from nigerian palm kernel
524 oil, Agricultural Engineering International: CIGR Journal (2007).

525 [54] F. Ma, L. D. Clements, M. A. Hanna, The effects of catalyst, free fatty acids,
526 and water on transesterification of beef tallow, Trans. ASAE 41 (5) (1998)
527 1261.

528 [55] H. J. Berchmans, S. Hirata, Biodiesel production from crude Jatropha curcas
529 L. seed oil with a high content of free fatty acids, Bioresour. Technol. 99 (6)
530 (2008) 1716–1721.

531 [56] C. H. Chen, W. H. Chen, C. M. J. Chang, S. M. Lai, C. H. Tu, Biodiesel pro-


532 duction from supercritical carbon dioxide extracted jatropha oil using sub-
533 critical hydrolysis and supercritical methylation, J. Supercrit. Fluids 52 (2)
534 (2010) 228–234.

535 [57] M. N. Varma, G. Madras, Synthesis of biodiesel from castor oil and linseed
536 oil in supercritical fluids, Ind. Eng. Chem. Res. 46 (1) (2007) 1–6.

537 [58] S. Saka, Y. Isayama, A new process for catalyst-free production of biodiesel
538 using supercritical methyl acetate, Fuel 88 (7) (2009) 1307–1313.

539 [59] J. S. Lee, S. Saka, Biodiesel production by heterogeneous catalysts and su-
540 percritical technologies, Bioresour. Technol. 101 (19) (2010) 7191–7200.

27
541 [60] D. Wen, H. Jiang, K. Zhang, Supercritical fluids technology for clean biofuel
542 production, Progress in Natural Science 19 (3) (2009) 273–284.

543 [61] M. Taylor, B. S. Atri, S. Minhas, P. Bisht, Developments in microwave


544 chemistry, Evaluserve Special Report (2005) 1–50.

545 [62] J. Geuens, J. M. Kremsner, B. A. Nebel, S. Schober, R. A. Dommisse,


546 M. Mittelbach, S. Tavernier, C. O. Kappe, B. U. W. Maes, Microwave-
547 assisted catalyst-free transesterification of triglycerides with 1-butanol under
548 supercritical conditions, Energy & Fuels 22 (1) (2007) 643–645.

549 [63] P. S. R. Fernandes, L. E. P. Borges, C. E. G. de Carvalho, R. O. M. A.


550 de Souza, Microwave assisted biodiesel production from trap grease, J. Braz.
551 Chem. Soc. 25 (9) (2014) 1730–1736.

552 [64] M. Nüchter, B. Ondruschka, A. Jungnickel, U. Müller, Organic processes


553 initiated by non-classical energy sources, J. Phys. Org. Chem. 13 (10) (2000)
554 579–586.

555 [65] T. M. Barnard, N. E. Leadbeater, M. B. Boucher, L. M. Stencel, B. A. Wil-


556 hite, Continuous-flow preparation of biodiesel using microwave heating, En-
557 ergy Fuels 21 (3) (2007) 1777–1781.

558 [66] A. A. Refaat, S. T. El Sheltawy, K. U. Sadek, Optimum reaction time, per-


559 formance and exhaust emissions of biodiesel produced by microwave irradi-
560 ation, IJEST 5 (3) (2008) 315–322.

561 [67] N. Kapilan, B. D. Baykov, A review on new methods used for the production
562 of biodiesel, Pet. Coal 56 (1) (2014) 62–73.

28
563 [68] D. Özçimen, S. Yücel, Novel methods in biodiesel production, in: M. A.
564 d. S. Bernardes (Ed.), Biofuel’s Engineering Process Technology, InTech,
565 2011, Ch. 16, pp. 353–384.

566 [69] D. Kumar, G. Kumar, C. P. Singh, Fast, easy ethanolysis of coconut oil for
567 biodiesel production assisted by ultrasonication, Ultrason. Sonochem. 17 (3)
568 (2010) 555–559.

569 [70] S. B. Lee, J. D. Lee, I. K. Hong, Ultrasonic energy effect on vegetable oil
570 based biodiesel synthetic process, Ind. Eng. Chem. Res. 17 (1) (2011) 138–
571 143.

572 [71] P. Cintas, S. Mantegna, E. C. Gaudino, G. Cravotto, A new pilot flow reac-
573 tor for high-intensity ultrasound irradiation. Application to the synthesis of
574 biodiesel, Ultrason. Sonochem. 17 (6) (2010) 985–989.

575 [72] A. S. Badday, A. Z. Abdullah, K. T. Lee, M. S. Khayoon, Intensification of


576 biodiesel production via ultrasonic-assisted process: a critical review on fun-
577 damentals and recent development, Renew. Sust. Energ. Rev. 16 (7) (2012)
578 4574–4587.

579 [73] M. C. Hsiao, C. C. Lin, Y. H. Chang, L. C. Chen, Ultrasonic mixing and


580 closed microwave irradiation-assisted transesterification of soybean oil, Fuel
581 89 (12) (2010) 3618–3622.

582 [74] X. Yin, H. Ma, Q. You, Z. Wang, J. Chang, Comparison of four different en-
583 hancing methods for preparing biodiesel through transesterification of sun-
584 flower oil, Appl. Energy 91 (1) (2012) 320–325.

29
585 [75] S. Sakthivel, S. Halder, P. D. Gupta, Influence of co-solvent on the produc-
586 tion of biodiesel in batch and continuous process, Int. J. Green Energy 10 (8)
587 (2013) 876–884.

588 [76] D. G. B. Boocock, S. K. Konar, V. Mao, H. Sidi, Fast one-phase oil-rich pro-
589 cesses for the preparation of vegetable oil methyl esters, Biomass Bioenerg.
590 11 (1) (1996) 43–50.

591 [77] R. Sawangkeaw, K. Bunyakiat, S. Ngamprasertsith, Effect of co-solvents


592 on production of biodiesel via transesterification in supercritical methanol,
593 Green Chem. 9 (6) (2007) 679–685.

594 [78] G. Knothe, Current perspectives on biodiesel, Information 13 (2002) 900–


595 903.

596 [79] S. V. Ghadge, H. Raheman, Biodiesel production from mahua (Madhuca


597 indica) oil having high free fatty acids, Biomass Bioenerg. 28 (6) (2005)
598 601–605.

599 [80] P. V. Rao, G. S. Rao, Production and characterization of jatropha oil methyl
600 ester, Int. J. Eng. Res. 2 (2) (2013) 141–145.

601 [81] M. Naik, L. C. Meher, S. N. Naik, L. M. Das, Production of biodiesel


602 from high free fatty acid Karanja (Pongamia pinnata) oil, Biomass Bioen-
603 erg. 32 (4) (2008) 354–357.

604 [82] S. E. Karatay, G. Dönmez, Microbial oil production from thermophile


605 cyanobacteria for biodiesel production, Appl. Energy 88 (11) (2011) 3632–
606 3635.

30
607 [83] Y. Chisti, J. Yan, Energy from algae: current status and future trends: algal
608 biofuels–a status report, Appl. Energy 88 (10) (2011) 3277–3279.

609 [84] G. Huang, F. Chen, D. Wei, X. Zhang, G. Chen, Biodiesel production by


610 microalgal biotechnology, Appl. energy 87 (1) (2010) 38–46.

611 [85] M. J. Montefrio, T. Xinwen, J. P. Obbard, Recovery and pre-treatment of


612 fats, oil and grease from grease interceptors for biodiesel production, Appl.
613 Energy 87 (10) (2010) 3155–3161.

614 [86] M. G. Kulkarni, A. K. Dalai, Waste cooking oil an economical source for
615 biodiesel: a review, Ind. Eng. Chem. Res. 45 (9) (2006) 2901–2913.

616 [87] M. J. Haas, A. J. McAloon, W. C. Yee, T. A. Foglia, A process model to


617 estimate biodiesel production costs, Bioresour. Technol. 97 (4) (2006) 671–
618 678.

619 [88] S. P. Singh, D. Singh, Biodiesel production through the use of different
620 sources and characterization of oils and their esters as the substitute of diesel:
621 a review, Renew. Sust. Energ. Rev. 14 (1) (2010) 200–216.

31

You might also like