Anda di halaman 1dari 13

Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

Contents lists available at SciVerse ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Review

The effects of catalysts in biodiesel production: A review


I.M. Atadashi, M.K. Aroua *, A.R. Abdul Aziz, N.M.N. Sulaiman
Chemical Engineering Department, Faculty of Engineering, University Malaya, 50603 Kuala Lumpur, Malaysia

A R T I C L E I N F O A B S T R A C T

Article history: Biodiesel fuel has shown great promise as an alternative to petro-diesel fuel. Biodiesel production is
Received 9 May 2011 widely conducted through transesterification reaction, catalyzed by homogeneous catalysts or
Accepted 14 July 2011 heterogeneous catalysts. The most notable catalyst used in producing biodiesel is the homogeneous
Available online 20 July 2012
alkaline catalyst such as NaOH, KOH, CH3ONa and CH3OK. The choice of these catalysts is due to their
higher kinetic reaction rates. However because of high cost of refined feedstocks and difficulties
Keywords: associated with use of homogeneous alkaline catalysts to transesterify low quality feedstocks for
Biodiesel
biodiesel production, development of various heterogeneous catalysts are now on the increase.
Production
Alkaline catalyst
Development of heterogeneous catalyst such as solid and enzymes catalysts could overcome most of the
Acid catalyst problems associated with homogeneous catalysts. Therefore this study critically analyzes the effects of
Solid catalyst different catalysts used for producing biodiesel using findings available in the open literature. Also, this
critical review could allow identification of research areas to explore and improve the catalysts
performance commonly employed in producing biodiesel fuel.
ß 2012 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2. Biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1. Techniques for biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2. Feedstocks for biodiesel production. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3. The effects of homogeneous catalyst in biodiesel production . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1. The effects of alkaline catalysts in biodiesel production. . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2. The effects of acid catalysts in biodiesel production. . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3. The effects of homogeneous catalysts on the yield of biodiesel fuel . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4. The effects of heterogeneous catalysts in biodiesel production. . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1. The effects of solid alkaline and acid catalysts in biodiesel production . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.1. The effects of solid alkaline catalysts in biodiesel production. . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.1.2. The effects of solid acid catalysts in biodiesel production. . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2. The effects of enzymes catalysts in biodiesel production . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.3. The effects of heterogeneous catalysts on the yield of biodiesel . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5. The advantages and disadvantages of homogeneous catalysts and heterogeneous catalysts. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
6. Conclusions and recommendations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

1. Introduction price hike, and the gradual depletion of fossil fuel reserves.
Therefore, to increase energy security for economic development,
Energy is the prime mover for socio-economic development. the need to search for an alternative source of energy such as
The World’s economic growth is affected by climatic change, fuel biodiesel is necessary [1,2]. Biodiesel is renewable, sustainable,
biodegradable, and emits low greenhouse gases [3,4]. As well, the
oxygen content of 11–15% in the molecular structure speed up the
* Corresponding author. Tel.: +60 3 79674615; fax: +60 3 79675319. combustion process in compression ignition engines and decreases
E-mail address: mk_aroua@um.edu.my (M.K. Aroua). pollutants such as soot, fine particles, and carbon monoxide (CO)

1226-086X/$ – see front matter ß 2012 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jiec.2012.07.009
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 15

O O cost of refined feedstocks, account to over 70% of the overall cost of


|| || biodiesel production [21]. As a result, different kinds of low quality
CH2 - O - C - R1 R4 - O - C - R1
feedstocks such as: waste cooking oils, used cooking oil, greases
| +
| O O CH2 - OH (yellow and brown), and non-edible oils have being investigated
| || || | [13,22]. The price of low quality feedstocks such as waste cooking
CH - O - C - R2 + 3 R4OH R4 - O - C - R2 + CH - OH oil is 2–3 times lower than refined oils. Nonetheless, the feedstocks
| (KOH) + | contains higher amount of FFAs and water contents. This features
| O O CH2 - OH makes their processing challenging [23,24]. Therefore to augment
| || ||
CH2 - O - C - R3 R4- O - C - R3 their processing difficulties, acid-catalyzed transesterification is
first employed to decrease the content of FFAs before performing
Triglyceride alcohol Alkyl Esters glycerin alkali-catalyzed transesterification [19]. Thus adopting two-step
transesterification technique could provide large biodiesel con-
Fig. 1. Transesterification reaction of triglycerides via alkaline catalyst.
version of up to 98% [25].
Recently heterogeneous catalysts such as solid catalysts and
[5,6]. Thus, biodiesel is a potential substitute to replace/supple- enzyme catalysts are employed to catalyze transesterification
ment petro-diesel fuel [7,8]. reaction for producing biodiesel. Heterogeneous catalysts offers
Biodiesel fuel is produced via transesterification of refined many advantages over homogeneous catalysts such as; simple
vegetable oil, waste cooking oil, and used frying oil using alkaline catalyst recovery, catalyst reusability, simple product purification,
catalysts [9–12] as shown in Fig. 1. The nature of catalyst employed less energy and water consumption, less added cost of purification,
during transesterification reaction is crucial in converting trigly- and simple glycerol recovery. Besides, most of the heterogeneous
cerides to biodiesel. As a result different catalysts have being catalysts used especially solid alkaline catalysts have provided
explored for converting triglycerides to biodiesel fuel. The catalysts high yields [26], though faced with problem of leaching [27]. Also,
usually employed to catalyze transesterification reaction are the stability of enzymes catalysts in non-aqueous media is
homogeneous catalysts and heterogeneous catalysts. Convention- significant to its excellent catalytic activity, this improves
ally, homogeneous alkaline catalysts such as NaOH, KOH, CH3ONa, transesterification and esterification during biodiesel production
and CH3OK are more often used in producing biodiesel [13]. The [28], and providing high biodiesel yield (95 wt%) [29]. However,
catalytic performance of these catalysts and their ability to the problem mostly associated with enzymes catalysts is the cost
perform under moderate conditions has led to their choice [14]. of the enzymes, but immobilization of the catalyst could mitigate
Among these homogeneous alkaline catalysts, CH3ONa is most the cost [30]. Therefore, to achieve biodiesel that is economically
active, providing biodiesel yield above 98 wt% in short reaction feasible, development of active and cheap catalysts for effective
time (30 min) [15,16]. However because of low price, industrial transesterification of different kinds of feedstocks is necessary
biodiesel production process mostly employs NaOH and KOH [14]. [31]. In this regards, this study extensively examined and reported
The process involving these catalysts needs high-quality feed- the effects of different catalysts in producing biodiesel fuel.
stocks, thus the free fatty acid (FFAs) level of the feedstocks should
not exceed 3 wt%, beyond which the reaction will not occur. In 2. Biodiesel production
addition, water content of the feedstocks is critical, as a result the
feedstocks used in alkali-catalyzed transesterification have to 2.1. Techniques for biodiesel production
anhydrous [14,17]. Thus, presence of water leads to hydrolysis of
oils to FFAs. Fig. 2 shows water hydrolysis of fats and oils to form Biodiesel is usually produced through different techniques such
free fatty acid. The FFAs react with alkaline catalysts to produce as direct/blends [32,33], microemulsion [34,35], pyrolysis [36,37]
soaps formation. Fig. 3 presents soaps formation in homogeneous and transesterification [38,39]. As stated earlier, alkali-catalyzed
alkali-catalyzed transesterification. Soaps formation consumes the transesterification is the most adopted technique for producing
catalyst, deactivates it and makes biodiesel purification process biodiesel, this method usually needs refined feedstocks containing
difficult [13,18]. Therefore, preparation of biodiesel by low quality less FFAs content otherwise it will result to much soaps formation.
feedstocks containing huge quantity of FFAs and water needs For alkali-catalyzed transesterification, if the feedstocks contains
sound technology [19]. However, high cost of refined feedstocks high amount of FFAs then it has undergo pretreatment steps before
result in high price of biodiesel compared to diesel fuel [15,20]. The transesterification [40]. Hideki et al. [14] and Ramadhas et al. [25]
recommended acid value of feedstocks to be less than 4.0 mg KOH/g
before performing alkaline transesterification. Although, Canakci
CH2 - O - CO - R1 CH2 - OH and Gerpen [41] and Mittelbach [42] stated that before alkali-
| |
catalyzed transesterification, the acid value of a feedstock has to be
| | O
| | || 2.0 mg KOH/g. Nonetheless, the use of heterogeneous catalysts in
CH - O - CO - R2 + H2O CH - O - CO - R2 + HO - C-R1 biodiesel production has reduced the effects of using low quality
| | feedstocks, especially enzymes catalysts that has the potential of
| | converting FFAs into biodiesel, besides high purity by-products [14].
| |
CH2 - O - CO - R3 CH2 - O - CO- R3
2.2. Feedstocks for biodiesel production
Triglyceride Water Diglyceride Free fatty acid
Biodiesel production is achieved via different kinds of feed-
Fig. 2. Water hydrolysis of fats and oils to form free fatty acid (FFAs). stocks. The nature of feedstock used is dependent on the
geographical position and climate of the place. For instance
O Europe employs sunflower and rapeseed oils, palm oil predomi-
|| + KOH R1 - COOK + H2O
nates in tropical countries, soybean in United States and canola oil
R1- C -OH
Free fatty acid Potassium hydroxide (Soap) Water in Canada [43]. Singh and Singh [44] reported the major feedstocks
employed in producing biodiesel are cotton seed, palm oil,
Fig. 3. Soap formation in homogeneous alkali-catalyzed transesterification. sunflower, soybean, canola, rapeseed, and Jatropha curcas.
16 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

Table 1 Table 3
Level of FFAs recommended for alkali-catalyzed transesterification. FFAs levels in feedstocks.

FFAs recommended (%) Reference Feedstocks FFA levels References

1 Demirbas [16] Trap grease 50–100% Sharma and Singh [3]


1 Chongkhong et al. [19] Refined vegetable oils <0.05% Davies [71]
<3 Ramadhas et al. [25] Finished greases 8.8–25.5% Canakci [41]
<0.5 Adam Khan [35] Crude soybean oil 0.4–0.7% Canakci [41]
<3 Canakci and Van Gerpan [41] Restaurant waste grease 0.7–41.8% Canakci [41]
<0.5 Zhang et al. [45] Waste palm oil >20% Balat and Balat [23]
<0.5 Martino et al. [47] Municipal sludge Up to 65% Bryan [58]
2 Sahoo et al. [48] Animal fat 5–30% Davies [71], Van Gerpen [69]
<1 Ma and Hanna [49] Trap grease 75–100% Davies [71], Van Gerpen [69]
<2 Huang and Chang [50] Use cooking oil 2–7% Van Gerpen [69]
0.5 Szczesna Antczak et al. [51] Waste oil 46.75% Nie [70]
1 Marchetti et al. [52]
<1 Tiwari et al. [53]

oils. Since biodiesel fuels from refined oils are costly when
compared to petro-diesel fuel. Besides, the application of such
Additionally, Zhang et al. [45] remarked that employing feedstocks feedstocks in biodiesel production could minimize competition
such as waste frying oils, non-edible oils, and animal fats, as between demand of edible vegetable oils and cost of biofuel [27]. In
feedstocks could be useful in producing biodiesel. Although, addition, application of vegetable oils as sources of biodiesel needs
Banerjee and Chakraborty [46] stated that FFAs contents in the great efforts to either develop more productive plant species with a
waste cooking oil should be kept within certain limit for reaction high yield of oil or to increase oilseeds’ production [72].
involving both acid- and alkali-catalyzed transesterification Further, many studies have reported microalgal oil as feed-
reactions. Otherwise these substances may cause severe difficul- stocks for producing biodiesel [73–78]. Demirbas [79] noted that
ties in refining of biodiesel products. Table 1 presents the microalgal oil is the only feedstock that can meet the global
recommended FFAs values for alkali-catalyzed transesterification demand for transport fuels. The author also reported that soon,
method. While Table 2 shows FFAs contents of different vegetable microalgal oil will become the most important feedstocks for
oils. In addition, Table 3 presents FFAs levels of most of the biofuel production. Singh and Gu [80] reported that microalgae
feedstocks used to produce biodiesel. The cost of feedstocks feedstocks are receiving great attention as sources of energy
decreases as FFAs content increases. In case of industrial biodiesel because of their quick growth potential coupled with reasonably
production, there is need for low-cost (high FFAs) feedstocks such high lipid, carbohydrate and nutrients contents. In addition,
as used cooking oils, waste cooking oils, and non-edible vegetable Demirbas [81] highlighted that microalgae possess much quicker
growth-rates than terrestrial crops. The author noted the per unit
Table 2 area yield of oil from algae is estimated to be from 20,000 to
Values of FFAs content of different vegetable oils. 80,000 l per acre, per year. In deed this is 7–31 times more than the
Vegetable oils FFA levels (%) Reference
next best crop, palm oil. Table 4 presents lipid content in the dry
biomass of various species of microalgae [82].
Polanga oil 22.0 Sahoo et al. [48]
Cottonseed oil 0.11 Sahoo et al. [48]
Tobacco oil 35.0 Veljkovic et al. [54]
Spent bleaching earth 24.1 Huang and Chang [50] Table 4
Rubber oil 17.0 Ramadhas et al. [25] Lipid content in the dry biomass of various species of microalgae [82].
Palm fatty acid distillate (PFAD) 93 Chongkong et al. [19]
Species Lipid content (% dryweight)
Pongamia pinnata 20 Naik et al. [55]
Palm oil 5.3 Balat and Balat [23] Botyococcus braunii 25–80
Rape seed oil 2.0 Balat and Balat [23] Chlamydomonas reinhardtii 21
Tall oil 100 Demirbas [56] Chlorella emersonii 28–32
Jatropha oil 14.0 Tiwari et al. [54], Bryan [57] Chlorella protothecoides 57.9
Tung oil 9.55 Bryan [57] Chlorella pyrenoidosa 46.7
Soybean soapstock >90 Bryan [57] Chlorella vulgaris 14–22
Ponagamia oil 0.64 Sureshkumar et al. [58] Crypthecodinium cohnii 20
Pongamia oil 2.69 Kaul et al. [59] Cylindrotheca sp. 16–37
Salvadora oil 1.76 Kaul et al. [59] Dunaliella primolecta 23
Moringa oleifera 2.9 Rashid et al. [60] Dunaliella salina 6
Karanja oil 2.53 Sharma and Singh [61] Dunaliella tertiolecta 35.6
Sorghum bug oil 10.5 Mariod et al. [62] Euglena gracilis 14–20
Mahua oil 21.0 Puhan et al. [63] Hormidium sp. 38
Mahua oil 19.0 Ghadge and Raheman [64] Isochrysis sp. 25–33
Madhuca indica 20.0 Kumari et al. [65] Monallanthus salina >20
Zanthoxylum bungeanum 45.5 Zhang and Jiang [66] Nannochloris sp. 30–50
Acid oil 59.3 Haas et al. [67] Nannochloropsis sp. 31–68
Karanja oil 2.53 Meher et al. [68] Neochloris oleoabundans 35–54
Trap grease 50–100% Sharma and Singh [3] Nitzschia sp. 45–47
Finished greases 8.8–25.5% Canakci [41] Phaeodactylum tricornutum 20–30
Crude soybean oil 0.4–0.7% Canakci [41] Pleurochrysis carterae 30–50
Restaurant waste grease 0.7–41.8% Canakci [41] Prymnesium parvum 22–38
Waste palm oil >20% Balat and Balat [23] Scenedesmus dimorphus 16–40
Municipal sludge Up to 65% Bryan [57] Scenedesmus obliquus 12–14
Animal fat 5–30% Van Gerpen [69] Schizochytrium sp. 50–77
Trap grease 75–100% Van Gerpen [69] Spirulina maxima 6–7
Use cooking oil 2–7% Van Gerpen [69] Spirulina platensis 4–9
Waste oil 46.75% Nie [70] Tetraselmis sueica 15–23
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 17

Base catalysed SN2 Transesterification

C
O Gly'

H-O-H CH -O-H
3 K-O-H
O

C
O Gly'
-
CH3-O

C O Gly'
H
CH3-O
O-CH3
H-O-H
O

C
HO-Gly'
CH3-O
HO-CH 3
-OH

Fig. 4. Mechanism of base-catalyzed transesterification reaction.

Also to highlight the significance of microalgae as feedstock for achieved 85.9 and 91.67 wt% yield of esters for NaOH and KOH and
biodiesel production, Tran et al. [83] have reviewed biodiesel 99.33 and 98.46 wt% yields of esters for CH3ONa and CH3OK
production via microalgal oil. The authors noted that catalysts such respectively. The authors recorded 98 wt% yields of esters for
as acid, base and zeolites catalysts can used in catalyzing methoxides after separation and purification steps were complet-
transesterification involving microalgal oil as feedstocks. ed. Further, less yields losses and negligible ester dissolution in
glycerol were observed with methoxides compared to hydroxides.
3. The effects of homogeneous catalyst in biodiesel production Umer et al. [90] used alkali catalyst to produce sunflower oil
methyl esters. They reported notable yield of 97.1 wt% at 60 8C. In
3.1. The effects of alkaline catalysts in biodiesel production addition, alkaline catalysts concentrations ranging from 0.5–1 wt%
could yield 94–99 wt% conversion of vegetable oils to alkyl esters.
Application of alkali-catalyzed transesterification reaction However, increase in catalyst concentration above 1 wt% does not
provide faster rate, nearly 4000 times faster than that catalyzed increase the conversion but could add to extra costs of production.
by the same amount of an acid catalyst [14]. Some of the alkaline Since it is essential to get rid of the catalyst from the products after
catalysts used for the transesterification reaction include among the reaction is completed [91,92].
others; NaOH [19,84], KOH [85,86], and sodium methoxide [16,84]. Chung [93] transesterified V. fordii and C. japonica seed oils with
Other alkaline catalysts include; sodium ethoxide [57], potassium methanol using alkaline catalysts (KOH, NaOH, and CH3ONa) to
methoxide [3,13], sodium propoxide [57], sodium butoxide [31] produce biodiesel. The authors noted that KOH provided higher
and carbonates [31,87], etc. Based on biodiesel yield, CH3ONa or catalytic activity to the seed oils in the reaction. The optimum
CH3OK are better and more suitable catalyst than NaOH and KOH. reaction conditions used were: 6:1 molar ratio of methanol to the
Thus, CH3ONa and CH3OK are more suitable due to their ability to seed oils, 1 wt% loading amount of catalyst, 65 8C reaction
dissociate into CH3O and Na+ and CH3O and K+ respectively. temperature, and reaction time of 3 h. The biodiesel contents of
Besides, the catalysts do not form water during transesterification the C. japonica and V. fordii seed oils under these reaction
reaction [3]. For these reason, alkaline catalyst is mostly preferred conditions were 97.7% and 96.1% on KOH catalyst. Figs. 5 and 6
in commercial production of biodiesel fuel [13]. Transesterification shows the effects of alkaline catalysts on feedstocks contain high
of refined oils with less than 0.5 wt% FFAs via chemical catalysts amount of water and FFAs [94]. The lesser the water and FFAs
could lead to high-quality biodiesel fuel with better yield within contents the better yield vice versa. In another study, Sahoo et al.
short time of 30–60 min [88]. Fig. 4 presents the mechanism of [48] employed alkaline transesterification after reducing the FFAs
base-catalyzed transesterification reaction [71]. Vicente et al. [89] value from 44 mg KOH/g a below 4 mg KOH/g through acid
have compared different basic catalysts (sodium hydroxide, catalyzed transesterification. The author found that 1.5 wt% of
potassium hydroxide, sodium methoxide and potassium methox- KOH was adequate to obtained maximum biodiesel yield.
ide) to produce biodiesel fuel using sunflower oil. The reactions Marchetti and Errazu [95] revealed that ethanol and sulfuric acid
were conducted at temperature of 65 8C, methanol to oil molar are suitable to carry out both direct esterification and transester-
ratio of 6:1 and basic catalyst by weight of vegetable oil of 1%. They ification reactions simultaneously. These processes could
18 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

Alkaline catalyst
separation time of 30 min. Also, Saifuddin and Chua [97]
optimized transesterification of used frying oil to ethyl ester
110
using microwave irradiation. They used a microwave oven
equipped with non-contact infrared continuous feedback tem-
100
perature system and magnetic stirrer to heat the oil and the
alcohol at 60 8C. Twenty five percent (25%) of an exit power of
90 750 W was used to irradiate the reaction mixture. The authors
Methyl esters (%)

experimented different concentrations of sodium methoxide


80 (0.3 wt% to 0.5 wt%) and achieved maximum conversion (87 wt%)
at 0.5 wt%. During transesterification process, both sodium
70 ethoxide and potassium hydroxide provided good conversions.
However, due simplicity in products phase separation, sodium
60 ethoxide was viewed as most promising catalyst for producing
biodiesel. Besides, microwave-assisted transesterification pro-
50 cess dramatically reduced the reaction time from 75 min to 4 min
at 60 8C, thus saving great time. Additionally during transester-
40 ification, irradiation times must be controlled and the levels of
0 5 10 15 20 25 30 35
radiation power should not be too high, to avoid destruction of
Water content (%)
organic molecules.
Fig. 5. Graph of methyl esters yield against water content in transesterification Moreover, to determine optimum operating parameters such as
reaction. catalyst concentration, Ferella et al. [98] studied transesterification
reaction of rapeseed oil for biodiesel production using response
surface methodology (RSM). They employed a 500 ml jacketed
Alkaline catakyst stirred (at 600 rpm) reactor tank. At optimum conditions of
temperature of 50 8C, KOH concentration of 0.6% (w/w), reaction
105 time of 90 min, and 60 methanol to KOH ratio by weight, large
amount of triglycerides, diglycerides and monoglycerides were
100
converted into biodiesel. As well, the final concentrations were
95 0.05% triglycerides, 0.09% diglycerides and 0.36% monoglycerides
and the triglyceride conversion was 98–99%.
Methyl esters (%)

90 The effects of catalysts such as NaOH on transesterification of


rapeseed oil under supercritical/subcritical conditions were also
85 studied by He et al. [99]. The authors employed little quantity of
NaOH as catalyst and achieved excellent biodiesel yield with no
80
soaps formation.
75
3.2. The effects of acid catalysts in biodiesel production
70
The most notable acids commonly employed in transester-
65 ification reaction include among others; sulfuric acid [13,19,100],
0 1 2 3 4 5 6
sulfonic acid [14], hydrochloric acid [16], organic sulfonic acid
FFA (%) [14], ferric sulfate [13], etc. Among these acids, hydrochloric acid,
Fig. 6. Graph of methyl esters yield against FFA in transesterification reaction. sulfonic acid and sulfuric acid are usually favored as catalysts for
the production of biodiesel. The catalyst and the alcohol are
vigorously mixed with a stirrer in a small reactor. The oil is first
effectively convert waste cooking oil containing high amount of charged into the biodiesel reactor and then the mixture of
FFAs ranging from 3% to 40% to biodiesel. The authors noted the catalyst/alcohol is fed into the oil. Brønsted acids preferably
FFAs content was reduced via esterification process from 10.684% sulfuric acid or sulfonic acid is used to catalyze the reaction.
to a value close to 0.54 wt%. Though, the final FFAs concentration Although the catalysts give high yield of biodiesel, but the reaction
was slightly more than the recommendable quantity. They rates are slow. The alcohol to oil molar ratio is the main factor
reported minimized soaps formation during alkali-catalyzed influencing the reaction. Therefore addition of excess alcohol
transesterification reaction. speeds up the reaction and favors the formation biodiesel
Further, to improve biodiesel production process, Refaat et al. products. The steps involve during acid-catalyzed transesterifica-
[96] studied microwave irradiation technique to produce tion are: (1) initial protonation of the acid to give an oxonium ion
biodiesel. They employed sunflower oil (used 3 times at a (2) the oxonium ion and an alcohol undergo an exchange reaction
cooking temperature of 130 8C) and methanol to oil ratio of 6:1 in to give the intermediate (3) and this in turn can lose a proton to
the presence of 1 wt% of potassium hydroxide at 65 8C. The become an ester. Reversibility of each of the above step is possible
authors used a normal pressure glass reactor 500 ml flask and but the equilibrium point of the reaction is displaced in the
reflux condenser. Using the microwave system, the vegetable oil presence of excess large alcohol, by allowing esterification to
was preheated to a desired temperature of 65 8C. The mixture of advance to completion [16]. Leung et al. [40] reported that
alcohol and catalyst then charged into the flask through the esterification by acid-catalysis makes the best use of the FFAs in
condenser. The power output adjusted to 500 W and under reflux the animal fats and vegetable oils and converts them into fatty
the mixture irradiated via different reaction times of 0.5, 1, 1.5, 2, acid alkyl esters. The authors noted that one-step esterification
2.5, 3 and 6 min. Using microwave irradiation technique, reaction pretreatment may not reduce the FFAs efficiently, if the acid value
time was reduced by 97% and the separation time reduced by 94%. of the oils or fats is very high. This is because of the high content of
They recorded biodiesel yield of 100% within 2 min and water produced during the reaction. In this case, addition of
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 19

Table 5
Reaction yield as function of the homogeneous catalyst weight.

Feedstock Catalyst Conc. (wt/v/v) Reaction time (h) Reaction temp (8C) Yield/conv. (w/w%) Molar ratios Reference

Waste tallow (chicken) H2SO4 1.25 24 50 99.01  0.71 1:30 [85]


Palm fatty acid H2SO4 – 2 70 99.6 7.2:1 [19]
Sunflower oil KOH 0.55 – 70 96 – [104]
Jojoba oil-wax Sodium methoxide 1 4 60 55 7.5:1 [105]
Brassica carinata KOH – – – 98.27 – [106]
Canola oil KOH 0.5 0.33 25 86.1 6:1 [107]
Jatropha curcas KOH 1 30 92 – [108]
Cottonseed oils Sodium hydroxide 0.5 1 55 77 – [109]
Roselle oil Potassium hydroxide 1.5 1 60 99.4 8:1 [110]
Rubber seed oil (Hevea brasiliensis) NaOH 1 1 60 84.46 6:1 [111]
Mahua oil (Madhuca indica) NaOH 1 2 60 92 6:1 [63]
Mahua oil H2SO4/KOH 1/0.7 1 60 98 6:1 [64]
Sunflower frying oil KOH 1 0.5 25 max 6:1 [112]
Tobacco seed oil NaOH 1.5 1.5 55 max 3:1 [113]
Rice bran oil Sulfuric acid 2 8 100 98 – [21]
Used frying oil KOH 1 2 60 72.5 12:1 [12]
Waste cooking oils KOH 0.75 0.33–2 30–50 88–90% 7:1–8:1 [11]
Jatropha oil H2SO4/KOH 0.25–1.5/0.5 2 60 90–95 6:1/9:1 [114]
Karanja oil KOH 1 3 65 97–98 6:1 [115]

mixture of alcohol and sulfuric acid into the oils or fats three times 4. The effects of heterogeneous catalysts in biodiesel
(three-step pre-esterification) is required. The time needed for production
this process is about 2 h and removal of water is necessary by a
separation funnel before adding the mixture into the oils or fats Despite the problems surrounding alkali-catalyzed transester-
for esterification again. Further, Palligarnai and Briggs [24] ification, the process is still favorable in producing biodiesel fuel.
reported sulfuric acid catalyzed transesterification to provide a The main reason is due to their faster kinetic rate and economic
few advantages over base-catalyzed technique such as one-step viability [116]. Recently several researches were conducted on
process as opposed to a two-step transesterification. The authors heterogeneous catalysts with the aim of finding solutions to
also reported that feedstock with a high FFAs content could be problems caused by using homogeneous catalysts in producing
easily handle, downstream separation is straightforward, and a biodiesel. As a result a good number of heterogeneous catalysts
high-quality glycerol byproduct is achievable. were explored and many of the catalysts have displayed very good
Additionally, Soriano Jr. et al. [101] studied transesterification catalytic performances [27]. Some of these catalysts include among
of canola oil to produce biodiesel via homogeneous Lewis acid others; oxides [117], hydrotalcides [118], zeolites [119], etc.
(AlCl3 and ZnCl2) as catalyst. The reaction occurred in a round Currently, majority of heterogeneous catalysts used in producing
bottom flask submerged in an oil bath equipped with a reflux biodiesel are either oxides of alkali or oxides of alkaline earth
condenser, temperature controller and a magnetic stirrer. The metals supported over large surface area [15]. Further, biodiesel is
authors reported use of variable parameters such as: reaction time commercially produced using heterogeneous catalyst through the
(6, 18, 24 h), methanol to oil molar ratio (6, 12, 24, 42 and 60), Esterfip-H process. This biodiesel process is commercialized by
reaction temperature (75, 110 8C). And tetrahydrofuran (THF) as Axens. The process needs neither catalyst recovery nor aqueous
co-solvent (1:1 methanol to THF by weight in runs with THF), and treatment steps, which are major bottlenecks from the current
catalyst (AlCl3 or ZnCl2). In all the runs, the catalyst amount was homogeneous catalytic processes. Additionally the Esterfip-H
kept at 5% based on the weight of oil. 1H NMR monitored process displays high biodiesel yields and directly produces salt-
converting canola oil into fatty acid methyl esters. Thus the best free glycerol at purities exceeding 98% compared to 80% glycerol
conditions with AlCl3 were 24:1 molar ratio at 110 8C and 18 h purity from homogeneous catalyzed process [120].
reaction time with THF as co-solvent provided a conversion of 98%. Thus, the effects of these catalysts are discussed as follows.
AlCl3 was far more active compare to ZnCl2 due to its higher acidity.
In another study, Cardoso et al. [102] discussed the effects of Lewis 4.1. The effects of solid alkaline and acid catalysts in biodiesel
acid on the transesterification process in producing biodiesel. The production
authors have introduced an inexpensive Lewis acid Tin(II) chloride
dihydrate (SnCl22H2O), and evaluated its potential as catalyst on Semwal et al. [121] reported that many studies on solid acidic
the ethanolysis of oleic acid of fats and vegetable oils. The catalytic catalysts for producing biodiesel were carried out, however lower
tests were carried out in triplicate with a molar ratio fatty acid: reaction rates and unfavorable side reactions have limited their
catalyst (100:1), reaction time of 2 h. Using these conditions, the uses. The authors noted that a good number of investigations on
process provided more than 90% biodiesel yield with a high basic heterogeneous catalysts were also conducted but their
selectivity of more than 93%. activity gets degraded in the presence of water. They stated that
acid–base catalysts are among the most promising catalysts to
3.3. The effects of homogeneous catalysts on the yield of biodiesel fuel employ in biodiesel production; this is because they can catalyze
both esterification and transesterification simultaneously. Further,
The nature of catalyst used plays a great role during Lee and Shiro [122] reported simultaneous esterification and
transformation of vegetable oil to biodiesel [103]. As a result, transesterification of waste cooking oil using solid catalyst ZnO–
catalyst is among the key factors determining the rate and yield of La2O3, which combines acid (ZnO) and base sites (La2O3). Although
biodiesel during biodiesel production process [17,46]. Thus, Table the process provided high conversion of 96% in 3 h, but like
5 presents reaction yield as function of the homogeneous catalyst zirconia, lanthanum is a rare and expensive metal, therefore cost of
weight. The data obtained shows that production of biodiesel via the catalyst would prohibit it high use in the production of
homogeneous catalyst could yield more than 99%. biodiesel. Further gelular resin catalysts having covalently bound
20 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

sulfonic groups are developed and employed to simultaneously reactivation procedures. As a result, the catalysts are suitable for
esterify and transesterify variety of feedstocks including beef producing biodiesel. In another study, Sharma et al. [128]
tallow and soybean oil of high acid number. The authors noted the remarked that the regeneration and reuse of CaO for many times
key factors influencing the feasibility and performance of solid (for instance, 20 runs) makes the catalyst most favorite among the
catalysts, are durability, catalytic activity and cost of production. oxides. These characteristics are useful from economic analysis.
Therefore the main challenges to R&D in using solid catalyst in Further, moderate conversion and yield could be obtained using
production of biodiesel are exploring the impacts of those chemical mixed oxides as catalyst therefore oxides of calcium and
properties of supports on the catalytic activity. And designing magnesium are preferred. Besides presence of some amount of
specifically tailored deactivation–prevention and/or renewal water/moisture in the reaction mixture has fewer effects on the
techniques for each catalyst and developing less expensive activity of CaO, this reduces pretreatment cost of feedstock and
supports. alcohol. However, Hsin et al. [129] reported that the high solubility
of CaO in methanol makes it a less attractive candidate for
4.1.1. The effects of solid alkaline catalysts in biodiesel production recyclable heterogeneous catalysis. Consequently the authors have
In recent times there is a great development in the preparation produced biodiesel fuel from soybean oil (SBO) and poultry fat
of solid alkaline catalysts for producing biodiesel. Solid alkaline using new calcium containing silicate mixed oxide-based hetero-
catalysts such as CaO provides many advantages for instance geneous catalysts. The catalysts are; PME-templated calcium
higher activity, long catalyst life times, and could run under only containing silicate mixed oxide catalysts (PMCS 1–9) and
moderate reaction conditions [123]. However, use of CaO in mesoporous calcium containing silicates (CS 1–9). The authors
biodiesel production presents slow reaction rate [31]. Antunes reported transesterification of SBO via PMCS-5 consisting of total
et al. [124] noted that like homogeneous catalysts, solid alkaline surface area of 150 m2/g to provide 100% yield within 2 h. This
catalysts present more catalytic activity than solid acid catalysts. catalyst could be recycled and reused for 8 times after regeneration
Meher et al. [115] examined methanolysis of karanja oil via solid by calcinations. They remarked that transesterification of soybean
basic catalysts. They remarked that increased in FFAs content of oil using CS-9 improved the reaction kinetics by a factor of 6. Also
karanja oil from 0.48 to 5.75% decreased the yield of biodiesel this catalyst could be recycled and reused for 3 times without any
produced decrease from 94.9 to 90.3 wt%. The authors achieved an decrease in reactivity for poultry fat and 9 times for soybean oil.
acid value of 0.36 mg KOH/g and an ester content of 98.6 wt%, after Additionally, Zabeti et al. [130] optimized biodiesel yields
purification of the esters. The properties of the esters produced met produced using transesterification of palm oil via CaO/Al2O3. They
both American and European standard specifications for biodiesel used a 150 ml glass jacketed reactor equipped with a water-cooled
fuels. condenser and a magnetic stirrer (1000 rpm) to perform the
Also, Liu et al. [125] have produced biodiesel fuel from experiments for 5 h. The estimated optimal reaction conditions
soybean oil using CaO as a solid base catalyst. The authors achieved were; reaction temperature of 65 8C, catalyst content of
reported that use of CaO as a catalyst could provide a numbers of 6 wt% and alcohol/oil molar ratio of 12:1, and achieved a
advantages such as: high activity, lengthen catalyst life and corresponding yield of 98.64 wt%. The authors reported that the
moderate condition of reaction. In a similarly study, Masato et al. values achieved from ICP-MS showed little leaching of CaO/Al2O3
[123] produced biodiesel fuel via solid base-catalyzed transes- active species into the reaction medium and the catalyst
terification at a reaction time of 1 h and obtained esters yield of successfully reused for two successive cycles. To further, examine
93 wt% for CaO, 12 wt% for Ca(OH)2, and 0% for CaCO3. The the potential of solid alkaline catalysts, Ilgen and Akin [131]
authors stated that CaO will probably provide good productivity studied different heterogeneous catalysts such as K2CO3/g-Al2O3,
as NaOH as well as easy recovery of products and environmental Na2CO3/g-Al2O3, LiOH/g-Al2O3, NaOH/g-Al2O3, g-Al2O3, and KOH/
benignity. They used CaO to transesterify waste cooking oil with g-Al2O3and reported FAME yield of 89.40%.
an acid value of 5.1 mg KOH/g. The yield of esters was above 99% In recent years hydrotalcites materials are receiving great
at a reaction time of 2 h. But a fraction of the catalyst transformed attention because they can be applied as precursors [132] and as
into calcium soap. Thus, solving this problem will entail removal catalyst [133]. Hydrotalcite-like compounds (HTlcs) are a category
of FFAs before transesterification reaction. The authors also of anionic and basic clays referred as layered double hydroxides
noted that due to neutralization reaction of the catalyst, (LDH) with the formula Mg6Al2(OH)16CO34H2O [15]. Thus
concentration of calcium in the produced biodiesel increased production of biodiesel was experimented using different hydro-
from 187 ppm to 3065 ppm. Conversely, Lim et al. [126] noted talcites such as activated Mg–Al hydrotalcites [133,134]. Zeng et al.
that transesterification reaction involving CaO needs longer [135] have examined activation of Mg–Al hydrotalcite catalysts for
reaction time. But the benefits gained from the process such as rape oil, and achieved 90.5% conversion. The production of
elimination of neutralization process, less waste generation and biodiesel from soybean oil via Mg–Al hydrotalcite was patented
prospect of catalyst reusability compensates the delay. The by Siano et al. [136]. The ratio of Mg/Al was found to affect the
authors also achieved biodiesel purity of 98.6  0.8% within 2.5 h. performance of the catalyst. And the catalytic activity was highest
Besides for CaO catalyst, the process generated 90.4% biodiesel at 3–8 ratio of Mg/Al. The authors remarked that the catalyst is best
yield compared to biodiesel yield of 45.5% for NaOH and 61.0% for suitable for oils containing up to 10,000 ppm quantity of water.
KOH catalysts. Further, it was observed that compared to biodiesel They achieved 92% conversion at a reaction condition: reaction
yield of 80% under anhydrous conditions using CaO, addition of time of 1 h, temperature of 180 8C, catalyst concentration of 5 wt%
2.03 wt% water into the reaction medium of 8 wt% catalysts, 12:1 and alcohol/oil weight ratio of 0.45. Similarly the catalytic
alcohol/oil molar ratio and at 3 h of reaction time, could provide performance of Mg–Al hydrotalcite for the transesterification of
biodiesel yield above 95%. Besides the catalyst active sites were vegetable oil to biodiesel was investigated by Di Serio et al. [137].
found not to leach and the activity of the catalyst was stable after 20 The authors noted that at higher reaction temperatures (215–
cycles of the reaction [27]. 225 8C); Mg–Al hydrotalcite displayed high catalytic activity. The
Similarly, Puna et al. [127] conducted an experimental study to transesterification was then conducted at methanol/oil weight
produce biodiesel via CaO and CaO modified with Li catalysts. Both ratio of 0.45 and catalyst concentration of 1 wt%, and esters yield of
catalysts showed good catalytic performances with high activity 94 wt% was achieved. Xie et al. [138] have experimented
and stability. In fact yields of biodiesel were higher than 92% in two methanolysis of soybean using Mg–Al hydrotalcite as solid base
consecutive reaction batches without expensive intermediate catalyst. The synthesis of the Mg–Al hydrotalcite was carried out
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 21

by employing co-precipitation technique of aqueous solution of higher concentration of super-basic sites. The authors reported
NaOH, Na2CO3, Al(NO3)39H2O, and Mg(NO3)26H2O under intense that biodiesel contents of 93.5 and 95.1 wt% were achieved at 60 8C
stirring. Increased in catalytic performance and improved basicity for zeolite X with and without sodium bentonite, respectively.
were observed by improving temperature of calcinations to 500 8C. Though, during the course of the reaction, the active species were
But the catalyst showed very low catalytic activity. And a observed to be leached out to the product [141]. Additionally,
conversion of 67% was recorded at an operating condition of Supamathanon et al. [142] have transesterified jatropha oil using
9 h, catalyst concentration 7.5 wt%, temperature of 65 8C, and heterogeneous catalysts consisting of potassium supported on NaY
alcohol/oil molar ratio of 15:1. Further, Georgogianni et al. [118] (K/NaY) with potassium loading of 4, 8 and 12 wt% to produce
have employed heterogeneous catalyst to produce biodiesel fuel biodiesel. The experiments were conducted in a 50 ml round-
using used frying oil. The authors conducted comparison between bottom flask equipped with a water-cooled condenser and heated
the catalytic activities of MgO supported on MCM-41 and Mg–Al with a water bath. They remarked that under reaction temperature
hydrotalcite base catalyst in the production of biodiesel fuel from of 65 8C, methanol to oil molar ratio of 16:1 and at a reaction time
soybean frying oil with alcohol (methanol). They reported the of 3 h, the catalyst with 12 wt% of potassium loading yielded
conversion of oil to biodiesel to be 97% and 85% respectively after optimum biodiesel yield of 73.4%.
24 h under the same reaction conditions. Further the hydrotalcite
showed greater activity which was attributed to its high basicity. 4.1.2. The effects of solid acid catalysts in biodiesel production
Even though Mg-MCM 41 has higher specific surface area of The replacement of homogenous catalysts by the solid acid
1289 m2/g compared to the low specific surface area of Mg–Al catalysts is useful for green chemistry [143]. Helwani et al. [15]
hydrotalcite of 82 m2/g. reported that compared to homogeneous acid, solid acid catalysts
Also, Wen et al. [139] have synthesized fatty acid methyl esters are preferred although their catalytic activities are low. This is due
from vegetable oil with methanol catalyzed by Li-doped magne- to fact that they contain a multiple sites with different strength of
sium oxide catalysts. A number of Li-doped MgO samples were Bronsted or Lewis acidity. Kathlene et al. [144] reported solid acid
prepared by the incipient wetness impregnation with the Li/Mg catalysts to have strong capacity to substitute liquid acids, thus
molar ratio in the range of 0.02–0.15. The catalyst weight was wiping out separation, corrosion and environmental problems. The
varied in the range of 3–15 wt% (methanol/oil molar ratio of 12, authors evaluated different solid catalysts for the production of
temperature of 338 K, residence time of 2 h). The results showed biodiesel from high FFAs feedstocks (waste cooking oil). The
that the formation of strong base sites was principally promoted by catalysts investigated include; zinc stearate/SiO2, MoO3/ZrO2,
adding Li. This results to an increase of fatty acid methyl esters WO3/ZrO2, WO3/ZrO2–Al2O3, MoO3/SiO2, TPA/ZrO2, and zinc
synthesis. In addition the catalyst with Li/Mg molar ratio of 0.08 ethanoate/SiO2. They stated that zinc stearate immobilized on
and calcined at 823 K exhibits the highest biodiesel yield of 93.9% silica gel was most active and stable. The catalyst recycled several
at 333 K with the molar ratio of methanol/soybean oil of 12 and the times at optimized conditions of reaction temperature of 470 K,
catalyst amount of 9 wt%. The use of the catalyst for the second and 1:18 molar ratio of oil to alcohol, and 3 wt% catalyst loading
third cycle decreased the fatty acid methyl esters yield from 79.3% without being deactivated. The authors recorded a maximum ester
and 71.0%, respectively. The biodiesel conversion decreases with content of 98 wt%. Catalyst recycling reduces the biodiesel
further increasing Li/Mg molar ratio above 0.08, which is most processing cost. In another study, Ngo et al. [145] have esterified
likely attributed to the separated lithium hydroxide formed by FFAs in greases (12–40 wt% FFA) using a diphenylammonium
excess Li ions and a concomitant decrease of BET values. The metal triflate acid catalyst immobilized onto two robust and highly
leaching from the Li-doped MgO catalysts was detected. The porous solid silica supports (MCM-48 and SBA-15). The authors
leaching from the catalyst and the agglomeration of crystallites evaluated the catalytic activities of the catalysts. The catalysts
caused the deactivation of the catalyst during the initial reaction were reported to be effective for the esterification of FFAs in
cycles. greases with a conversion to biodiesel of 95–99%, resulting in a
Furthermore, Suppes et al. [119] have transesterified soybean pretreated grease with a final FFAs content of <1 wt%. Also, Furuta
oil with zeolites (ETS-10 zeolite, NaX faujasite zeolite) and metal et al. [146] studied biodiesel fuel production via transesterification
catalysts. The transesterification reaction was performed at 6:1 using soybean oil and methanol with solid superacid catalysts
molar ratio of alcohol and temperatures of 60, 120, and 150 8C. The sulfated tin oxide (SO4/SnO2) and zirconium oxide (SO4/ZrO2) and
conversion to esters increases from 60 to 150 8C with an average tungstated zirconia (WO3/ZrO2) at 200–300 8C. And conducted
conversion of 90% achieved at 125 8C. Thus, ETS-10 gave better esterification of n-octanoic acid with methanol at 175–200 8C in
conversions of triglycerides than zeolite-X type catalysts. This was fixed bed reactor under atmospheric pressure. The authors noted
attributed to the higher activity of ETS-10 zeolites and larger pore that out of the catalysts prepared, tungstated zirconia–alumina
structures that improved intra-particle diffusion. The catalyst catalyst (WZA) showed high performance, yielding over 90%
activity was not affected after reused. However, FFAs loadings in conversion for esterification processes. Although detail analysis for
excess of 25% quench the catalyst activity. Meyer et al. [140] the acidity of WZA has not yet been determine.
conducted experimental studies on heterogeneous catalysts to Besides, Kitiyanan et al. [147] studied transesterification of
produce biodiesel. The reaction occurred at temperature of 90 8C palm and coconut oils using solid catalysts such as KNO3/ZrO2,
and a methanol to triglyceride ratio of 9:1. The authors stated that KNO3/KL zeolite, SO42 /ZrO2, SO42 /SnO2, and ZrO2, ZnO. The
among the studied catalysts the potassium exchanged low silica authors noted that transesterification of crude palm kernel oil
zeolite of the faujasite type (K-LSX) has shown the best results and using SO42 /SnO2 and SO4/ZrO2catalysts provided highest yield of
reached ester formation of over 90 wt% after 1 h. The result methyl esters (90.3 wt%). The purities of the esters were 95.4 wt%
obtained was comparable to those of homogeneous catalysts such for SO42 /SnO2 and 95.8 wt% for SO42 /ZrO2, respectively. Howev-
as alkali metal hydroxides. Another study conducted by Marı́a et al. er, for ZnO the highest content of the esters was 98.9 wt%. In
[141] examined different zeolite catalysts such as NaX for the addition, owing to the availability of enough acid site strength,
parent zeolite in sodium form: 0.3MNaX, 1MNaX, 3MNaX, solid acid catalysts for example sulfated zirconia, tungstated
3MNaXB, and samples from mordenite and beta zeolite were zirconia and Nafion-NR50 were selected to catalyze transester-
named as 3NaM and 3Nab for the transesterification of sunflower ification [147,148]. However, due to its acid strength, the
oil with alcohol (methanol). Hence, zeolite X provided greater selectivity of Nafion-NR50 towards biodiesel and glycerol produc-
activity compared to other zeolites (mordenite and beta) due to its tion was higher, compared to both sulfated zirconia and tungstated
22 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

zirconia [149]. Nonetheless, Chai et al. [150] remarked that the consecutive additions of 1:3 molar equivalent of methanol. And
major drawbacks of Nafion as catalyst to produce biodiesel, is it then developed a three-step methanolysis by which over 95% of
lower catalytic activity and high cost compared to liquid acids. the oil triacylglycerols (TAG) transformed to their corresponding
Further, application of amberlystTM BD20 catalyst has provided esters.
more impetus for the development of biodiesel fuel. The Additionally, Shah et al. [30] have prepared biodiesel from
amberlystTM BD20 process can effectively convert any feedstock jatropha oil using lipase catalyst. The authors have screened
containing 0.5–100% FFAs to biodiesel. Further amberlystTM BD20 pancreas porcine, candida rugosa, Chromobacterium viscosum and
presents fast solid esterification catalyst. Besides high biodiesel in a solvent-free system for the production of biodiesel. They
yield can be obtained with improved biodiesel and glycerol purity employed a screw-capped vial and jatropha seed oil (0.5 g) and
[151]. Additionally the search for a suitable solid catalyst has lead ethanol were taken in the ratio of 1:4 (mol mol 1). Also 50 mg of
to the development of vanadium phosphate catalyst by Di Serio enzyme preparation (tuned or immobilized) was added and
et al. [152] for biodiesel production. The authors reported 2–4 m2/g incubated at 40 8C with constant shaking at 200 rpm. The
as specific surface area of the catalyst. However, despite the surface immobilization of lipase (Chromobacterium viscosum) on Celite-
area of the catalyst is low, the catalyst was found to be active in 545 improved yield of esters from 62% to 71% by free tuned enzyme
producing biodiesel from soybean oil. They reported ester yield of preparation with a process time of 8 h at 40 8C. Additionally to
80% at 150 8C and observed catalysts deactivation at elevated explore more information on the use of enzymes for the production
temperatures owing to reduction of V5+ to V3+ with the alcohol of biodiesel, Tan et al. [28] reviewed biodiesel production using
(methanol). Although regenerating process of the catalyst after use immobilized lipase. The immobilized lipase as biocatalyst draws
is simple and could be achieved without complexities. great interest because that process is environmentally friendly.
Also, zeolites are used as a catalyst for esterification and as a The authors noted different techniques for lipase immobilization,
support material for transesterification catalysts [2,142]. Zeolites such as covalent bonding, cross-linking, encapsulation, entrap-
are microporous crystalline solids with well defined structures and ment and adsorption. Lipase immobilization technique is com-
that they contain aluminum, silicon and oxygen in their framework monly used to increase the stability of lipase in biodiesel
and cations. As catalysts, unlike the compositionally equivalent production. They stated that, for biodiesel (fatty acid methyl
amorphous catalysts, zeolites demonstrate substantial acid esters) preparation, at least a stoichiometric amount of methanol is
activity and shape selective features [153–155]. Chung et al. needed for the complete conversion of triglycerides to their
[153] employed different Si/Al molar ratio to remove FFAs in waste resultant fatty acid methyl ester. But, methanolysis is reduced
frying oil by esterification with methanol using different zeolite considerably by adding >1/2 molar equivalent of methanol at the
catalysts. The catalysts used include mordenite (MOR), faujasite commencement of the enzymatic process. Usually, the polar short
(FAU), beta (BEA) zeolites, ZSM-5 (MFI), and silicalite. The pore chain alcohols causes inactivation of enzymes and this is the major
structure and the acidic properties of the zeolites were particular obstacle for the enzymatic-transesterification reaction. Therefore
useful in the removal of FFAs. These properties influenced the to overcome this difficulty one of these options is selected: acyl
catalytic activity in FFAs removal. High conversion of FFAs was acceptor alterations, solvent engineering and methanol stepwise
comparatively induced by strong acid sites of zeolites. The MFI addition. The authors reported that biodiesel yield of 97 wt% was
zeolite induced an improvement of the FFAs removal efficiency by obtained after 24 h at temperature of 50 8C with a reaction mixture
cracking to the FFAs in its pore structure owing to its constricted containing 13.5% methanol, 32.5% t-butanol, 54% oil and
pore mouth. Converting FFAs on HMFI and HMOR zeolites provided 0.017 g enzyme (g oil) 1. With the same mixture, a 95% biodiesel
80% at a reaction temperature of 60 8C. In addition, HMOR zeolites yield was achieved using a one step fixed bed continuous reactor
showed almost a similar conversion of FFAs with different Si/Al with a flow rate of 9.6 ml h 1 (g enzyme) 1. The authors concluded
molar ratio, but with decreased in acidity. Decreased in acid that low cost of immobilized Candida sp. 99–125 lipases is rather
strength of the zeolites lowered the catalytic performance for FFAs competitive for industrial use [28]. More so, Hideki et al. [14] noted
removal [156]. use of enzymatic-catalyzed transesterification to avoid problems
associated with homogeneous catalysts. Therefore the production
4.2. The effects of enzymes catalysts in biodiesel production processes are compared in Figs. 7 and 8. The authors reported
conversion of high FFAs feedstocks to biodiesel using immobilized
Transesterification reaction can be catalyzed with enzymes antarctica lipase (Novozym-435) with ease of separation process.
catalysts such as Candida antarctica lipase [28], Pseudomonas
cepacia [157], Candida sp. 99–125 [28,158], Pseudomonas fluor- 4.3. The effects of heterogeneous catalysts on the yield of biodiesel
escens [157], Rhizomucor miehei and Chromobacterium viscosum
[20,159] and Rhizopus oryzae lipase [160]. Casimir et al. [161] noted As earlier stated great efforts have being made by several
that biodiesel can be excellently produced via enzymatic- researchers and industries to explore and exploit use of heteroge-
transesterification reactions involving lipases. They suggested neous catalysts in the production of biodiesel. Some of the reasons
that large quantities of lipases can be produced using recombinant for the recent growth and development of heterogeneous catalysts
DNA technology. The authors believed that application of include among others: biodiesel yield of 98 wt% and simplicity in
immobilized lipase may reduce the overall cost of biodiesel catalyst separation process [130], high-purity by-products, less
production and lower downstream processing problems. Besides, cost of separation and low energy consumption [31]. Table 6
enzymatic approach is environmentally friendly [161]. Watanabe presents reaction yield as function of the heterogeneous catalyst
et al. [29] explored use of immobilized Candida antarctica lipase for weight.
continuous production of biodiesel fuel from vegetable oil. The
transesterification of vegetable oil was conducted using 4% 5. The advantages and disadvantages of homogeneous
immobilized Candida lipase as a catalyst at 30 8C in a 20- or 50- catalysts and heterogeneous catalysts
ml screw-capped vessel, shaking at 130 oscillations/min. The
authors noted the activity of Candida antarctica lipase was not The choice of catalyst to use in the production of economically
affected in a mixture of vegetable oil and more than 1:2 molar viable biodiesel fuel is one of the most critical issues that are being
equivalent of methanol against the total fatty acids. They discussed by futurists and industries [117]. As discussed earlier,
discovered inactivation of the lipase to be eradicated via three NaOH and KOH are mostly preferred because they are cheaper than
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 23

Alcohol for recycling


Vegetable
Crude
oils/Animal
biodiesel
fats Centrifuge/settler
Transesterification for separation of Neutralization Distillation/ Drying
Alcohol
Reaction crude biodiesel process/water Evaporation Facility
products washing
Alkali
Waste
Crude water
glycerin/ Mineral acid (H3PO4) Finished
impurities product
Pure glycerin for (biodiesel)
Mineral acid (H2SO4) Glycerin purification
Storage/recycling/sale

Free Fatty Acid (FFAs)

Fig. 7. Traditional alkali biodiesel production.

Fats and oils


Separation/purification Upper layer
Alcohol Finished biodiesel
Transesterification reaction of transesterified
product for
products
Enzymes consumption

Lower layer

Pure glycerin
Glycerin purification
for sale

Fig. 8. Enzymatic biodiesel production.

alkoxides, but are less active [17]. Leung and Guo [38] revealed that washings are needed to remove the catalysts. Zhang et al. [45]
industrially, separation of esters is much easier when KOH catalyst conducted economic analysis of four continuous processes to
is used compared to NaOH or CH3ONa. Because potassium soap produce biodiesel such as alkali- and acid-catalyzed processes,
formation is much softer and does not sink into glycerol phase. with virgin vegetable oil and waste cooking oil. The authors
Owing to this reason, KOH is most often employed to produce revealed that acid-catalyzed transesterification with waste cook-
biodiesel from waste recycled feedstocks [38]. Saka and Isayama ing oil were more economically viable. The process gave lesser
[170] reported that even though the alkaline catalyst technique has total production cost, a more attractive after-tax rate of return and
the advantage of using moderate reaction conditions, several water a less biodiesel break-even price.

Table 6
Reaction yield as function of the heterogeneous catalyst weight.

Feedstock Catalyst Conc (wt) Reaction time (h) Reaction temp (8C) Yield (w/w%) Molar ratios Reference

Soybean oil CaO 25 1 – 93 – [121]


Vegetable oil M(3-hydroxy-2-methyl-4-pyrone)2(H2O)2 3 60 93 – [162]
Sunflower oil CaO/SBA-14 5 160 95 12 [163]
Sunflower Zeolite X 4.2 – 60 95.1 – [164]
Soybean oil ETS-10 zeolite 0.03 24 125 90 – [119]
Palm oil CaO/Al2O3 3.5 5 65 98.64 12:1 [26]
Waste bleaching earths Rhizopus oryzae 96 35 55 1:4 [160]
Soybean Na/NaOH/g-Al2O3 1 2 60 96 9:1 [165]
Crude palm kernel oil SO42 /ZrO2 1 1 200 90.3 6:1 [144]
Waste cooking oil ZS/Si 3 10 200 98 1:18 [166]
Soybean oil Lipase 0.9 6.3 36.5 92.2 3.4:1 [159]
Palm oil CaO/Al2O3 3.5 5 65 95% 12:1 [130]
Canola oil Nano-g-Al2O3 8 65 97.7  2.1 15:1 [167]
Mixture of soybean and Candida lipase 4 24 30 93 1:3 [29]
rapeseed oils
Jatropha oil Chromobacterium viscosum lipase 5 8 40 92 1:4 [30]
Soybean oil Pseudomonas cepacia lipase 47.5 1 35 67 mol% 1:7.5 [168]
Sunflower oil WO3/ZrO2 3 5 200 97 20:1 [169]
24 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

Table 7
Advantages and disadvantages of homogeneous catalysts and heterogeneous catalysts.

Type of catalyst Example Merits Demerits Ref.

Homogeneous catalyst
Alkaline catalysts: NaOH, KOH, CH3ONa, CH3OK Less corrosive, high reaction rate Formation of saponified product, [13,17,98,153,
emulsion formation, high water 170,177–179]
and energy consumption, huge,
high wastewater discharges,
high purification cost. Feedstocks
are limited to 0.5 wt FFAs,
not recycle
Acid catalysts: H2SO4 Zero soap formation, the catalyst More waste as a result of [31,177,180]
can be used to catalyze both neutralization, recycling difficulty,
esterification and high purification cost, energy
transesterification consuming, low reaction rates
simultaneously
Heterogeneous catalysts
Solid alkaline catalysts MgO, CaO, ZnO KOH/NaY, Environmentally friendly, easily Leaching effects, catalysts [15,27,47,150,
and solid acid catalyst CaO/MgO, Al2O3–SnO, KOH/K2CO3, ecycle, less discharges, less preparation is complicated 152,162,172,
Al2O3–ZnO, Ca(NO3)2/Al2O3, separation difficulty, high and expense relatively slow rates 174,175,181,
CaO/Al2O3, KOH/Al2O3, Al2O3/KI purity glycerol, lower cost of separation 182]
Sr(NO3)2/ZnO, ZrO2/SO42 TiO2/SO4 2 . Insignificant leaching of CaO/Al2O3 [26,130]
ETS-10 zeolite, zeolite HY, and zeolite X
Enzymes catalysts: Candida antarctica B lipase, Rhizomucor Zero saponification products Catalysts inhibition by water [14,40,161,
meihei lipase, candida rugosa Pseudonas nonpolluting, easily separable, 180,183]
cepacia, M. meihei (Lypozyme), M. meihei lesser separation cost, high
(Lypozyme IM60), Aspergillus niger. purity glycerol and biodiesel
products, environmentally benign
P. fluorescens, R. Oryzae Simple glycerol recovery High cost of enzymes [29,30]

In contrast both alkali- and acid catalyzed transesterification advantage of employing enzyme catalysts is lack of soaps
are associated with a number of problems. Madras et al. [171] formation. They reported application of insoluble solid catalysts
reported that alkali-catalyzed transesterification present high cost (immobilized enzymes) to speed up catalyst removal from both
of biodiesel production and large energy consumption during glycerol and alkyl esters. Zabeti et al. [27] showed solid-catalyzed
down-stream biodiesel refining process. Ferella et al. [98] stated transesterification reaction to yield biodiesel products with no
that potassium hydroxide produces soaps formation by neutraliz- catalyst impurities. Therefore absence of leaching improves
ing FFAs. Additionally, due to the tendencies to produce soaps separation process and lowers the cost of final refining. Martino
formation and mono- and di-glycerides during transesterification, et al. [47] noted that heterogeneous catalysts can be separated
direct use of sodium and potassium alkylates as catalysts is raising more easily, besides low processing costs and zero waste streams.
serious issues in the industries. The soaps formations cause Hameed et al. [174] stated that heterogeneous catalysts are
formation of gels and makes separation and purification of searched over time to achieve environmentally benign and
biodiesel hard. Also, Cardoso et al. [102] reported that the economically viable biodiesel production and purification pro-
problems associated with H2SO4 catalyst include much reactor cesses. However, issues such as leaching of catalysts especially
corrosion and huge wastewater discharges resulting from the when solid alkaline catalysts such as CaO, MgO, La2O3, ZnO, SrO,
neutralization of mineral acid. Another investigation conducted by Li-promoted CaO, CaCO3, Ba(OH)2, KxX/Al2O3 (X-halide ion or
Helwani et al. [15] revealed that application of homogeneous other mono/di-valent anion), Zinc aluminates, Metal salts of
catalysts in biodiesel production presents difficulty in biodiesel amino acids, Mg–Al hydrocalcites and K- and Cs-exchanged
product separation and makes recovery of used catalyst cumber- zeolites are used in producing biodiesel creates great deal of
some. The authors reported that biodiesel production via concern during separation process. The leaching of active sites of
homogeneous alkaline catalysts needs multi-steps of production solid alkaline catalysts into liquids results from its characteristic
and purification, since such catalysts do not tolerate presence of changed turning it partly ‘‘homogeneous’’. This changed in the
moisture or FFAs. In addition, there are many notable issues properties of the catalysts affects biodiesel quality, makes catalyst
surrounding the existing transesterification processes such as; separation and biodiesel purification difficult, thus bringing extra
difficulties in processing low quality feedstocks and refining of cost of production [15]. Mariscal et al. [172] reported significant
transesterified products [40]. leaching of potassium during transesterification of triglyceride via
Further heterogeneous catalysts provide many advantages K/c-Al2O3 catalysts. The authors reported 99 wt% biodiesel yield
compared to homogeneous catalysts such as: catalyst re-usability, which was attributed to homogeneous contribution from active
less separation and purification difficulties, high purity glycerol basic species dissolved in the methanol. In addition, the
(above 99%), and catalyst recoverability [117,172]. Additionally homogeneity of this catalyst complicates its separation process.
no neutralization step is required [173]. Thus, in biodiesel Also leaching of solid acid catalyst (sulfate species) was reported
production, the overall economy can be improved through the to restrict its reusability [175]. Moreover, a study carried out by
use or sale of by-product, glycerol [117]. Mariscal et al. [172] Lee et al. [176] remarked that leaching of base active species
reported that heterogeneous catalysts are less corrosive, leading caused great negative effects on the degree of purity of biodiesel
to safer, less costly, and more environmentally friendly opera- and glycerol. Another investigation conducted through optimiza-
tions. Watanabe et al. [29] reported enzymatic-transesterification tion of catalytic activity of CaO/Al2O3 for producing biodiesel with
reaction to overcome problems associated with chemical pro- response surface methodology showed insignificant leaching of
cesses. Palligarnai and Briggs [24] found glycerol from enzymatic- catalyst into the reaction medium [26,130]. Table 7 summarizes
catalyzed transesterification not to show negative effects on the merits and demerits of homogeneous catalysts and
biodiesel fuel property. The authors stated that the major heterogeneous catalysts for biodiesel production.
I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26 25

6. Conclusions and recommendations [21] S. Zullaikah, C. Lai, S.R. Vali, Y. Ju, Bioresource Technology 96 (2005) 1889.
[22] K.-W. Lee, J.X. Yu, J.H. Mei, L. Yan, Y.-W. Kim, K.-W. Chung, Journal of Industrial
and Engineering Chemistry 13 (2007) 799.
Based on the foregoing the following conclusions and recom- [23] M. Balat, H. Balat, Applied Energy 87 (2010) 1815.
mendations were made: [24] T.V. Palligarnai, M. Briggs, Journal of Industrial Microbiology & Biotechnology 35
(2008) 421.
[25] A.S. Ramadhas, S. Jayaraj, C. Muraleedharan, Fuel 84 (2005) 335.
1. Use of high FFAs feedstocks to produce biodiesel is more [26] M. Zabeti, W.M.A.W. Daud, M.K. Aroua, Applied Catalysis A: General 366 (2009)
economically viable, though it has great negative effects on the 154.
[27] M. Zabeti, W.M.A.W. Daud, M.K. Aroua, Fuel Processing Technology 90 (2009)
catalyst. Besides refining of crude biodiesel produced from low 770.
quality feedstock with alkaline catalysts is complicated. [28] T. Tan, J. Lu, K. Nie, L. Deng, F. Wang, Biotechnology Advances 28 (2010) 628.
2. It was found that acid-catalyzed transesterification reaction [29] Y. Watanabe, Y. Shimada, A. Sugihara, H. Noda, F. Hideki, Y. Tominaga, JAOCS 77
(4) (2000).
needs higher alcohol to oil molar ratios. In addition the reaction
[30] S. Shah, S. Sharma, M.N. Gupta, Energy & Fuels 18 (2004) 154.
was reported to cause large corrosion effects on the production [31] L.C. Thiam, S. Bhatia, Bioresource Technology 99 (2008) 7911.
facilities. [32] A.L. Boehman, Fuel Processing Technology 86 (2005) 1057.
3. It observed that biodiesel can be conveniently produced from [33] A. Keskin, M. Gürü, D. Altiparmakc, K. Aydin, Renewable Energy 33 (2008) 553.
[34] A.S. Ramadhas, S. Jayaraj, C. Muraleedharan, Renewable Energy 29 (2004) 727.
low quality feedstocks using heterogeneous catalysts. Especial- [35] A. Khan, Evaluating Biodiesel Catalysts, 2007, p. 25, www.eptq.com.
ly, enzymatic-catalyzed transesterification reaction that was [36] L. Brennan, P. Owende, Renewable & Sustainable Energy Reviews 14 (2010) 557.
discovered to converts FFAs to biodiesel fuel. [37] S.N. Naik, V.V. Goud, P.K. Rout, A.K. Dalai, Renewable & Sustainable Energy
Reviews 14 (2010) 578.
4. The conversion of FFAs to biodiesel was found to increase the [38] D.Y.C. Leung, Y. Guo, Fuel Processing Technology 87 (2006) 883.
yield of biodiesel and simplify the separation of catalyst form [39] A. Salahi, M. Aoobbasi, T. Mohammadi, Desalination 251 (2010) 153.
crude biodiesel mixture. [40] D.Y.C. Leung, X. Wu, M.K.H. Leung, Applied Energy 87 (2010) 1083.
[41] M. Canakci, J.V. Gerpen, American Society of Agricultural Engineers 42 (1999)
5. To make biodiesel production process cost effective, the 1203.
chemistry of heterogeneous catalysts such as solid alkaline [42] M. Mittelbach, B. Pokits, A. Silberholz, ASAE Publication Nashville, TN, USA, 1992,
catalysts needs to be thoroughly explored, developed and use p. 74.
[43] P. Cao, M.A. Dubé, A.Y. Tremblay, Biomass & Bioenergy 32 (2008) 1028.
consistently in producing biodiesel. Further the catalysts were [44] S.P. Singh, D. Singh, Renewable & Sustainable Energy Reviews 14 (2010) 200.
found provide biodiesel yield similar those obtained from [45] Y. Zhang, M.A. Dubé, D.D. McLean, M. Kates, Bioresource Technology 89 (2003) 1.
homogeneous alkaline catalysts. [46] A. Banerjee, R. Chakraborty, Resources Conservation and Recycling 53 (2009)
490.
6. Tremendous effort should be made to exploit ways of improving
[47] D. Martino, R. Tesser, L. Pengmei, E. Santacesaria, Energy & Fuels 22 (2008) 207.
the kinetic rate of CaO catalysts, since its catalytic activity is [48] P.K. Sahoo, L.M. Das, M.K.G. Babu, S.N. Naik, Fuel 86 (2007) 448.
high, its catalyst life time is long and it operates under only mild [49] F. Ma, L.D. Clements, M.A. Hanna, Bioresource Technology 70 (1999) 1.
reaction conditions. [50] Y. Huang, J.I. Chang, Renewable Energy 35 (2010) 269.
[51] M. Szczesna Antczak, A. Kubiak, T. Antcza, S. Bielecki, Renewable Energy 34
7. Development of enzymes catalyst as a more environmentally (2009) 1185.
benign process is necessary due to the prevailing environmental [52] J.M. Marchetti, V.U. Miguel, A.F. Errazu, Renewable & Sustainable Energy
needs. Besides the process is potentially sound providing high- Reviews 11 (2007) 1300.
[53] A.K. Tiwari, A. Kumar, H. Raheman, Biomass & Bioenergy 31 (2007) 569.
quality biodiesel fuel that can compete favorably with petro- [54] V.B. Veljkovic, S.H. Lakicevic, O.S. Stamenkovic, Z.B. Todorovic, M.L. Lazic, Fuel 85
diesel fuel. (2006) 2671.
[55] M. Naik, L.C. Meher, S.N. Naik, L.M. Das, Biomass & Bioenergy (2007), http://
dx.doi.org/10.1016/j.biombioe.2007.10.006.
[56] A. Demirbas, Energy Sources, Part A 30 (2008) 1896.
References [57] R.M. Bryan, In Vitro Cellular & Developmental Biology: Plant 45 (2009) 229.
[58] K. Sureshkumara, R. Velrajb, R. Ganesana, Renewable Energy 33 (2008) 2294.
[59] S. Kaul, R.C. Saxena, A. Kumar, M.S. Negi, A.K. Bhatnagar, H.B. Goyal, et al.
[1] K.K. Oh, Y.S. Kim, H.H. Yoon, B.S. Tae, Journal of Industrial and Engineering Corrosion Fuel Processing Technology 88 (2007) 303.
Chemistry 8 (1) (2002) 64. [60] U. Rashid, F. Anwar, B.R. Moser, G. Knothe, Bioresource Technology 99 (2008)
[2] B.-H. Uma, Y.-S. Kim, Review: Journal of Industrial and Engineering Chemistry 15 8175.
(2009) 1. [61] Y.C. Sharma, B. Singh, Fuel (2007), http://dx.doi.org/10.1016/j.fuel.2007.08.001.
[3] Y.C. Sharma, B. Singh, Renewable & Sustainable Energy Reviews 13 (2009) 1646. [62] A. Mariod, S. Klupsch, H. Hussein, B. Ondruschka, Energy & Fuels 20 (2006)
[4] S.B. Lee, K.H. Han, J.D. Lee, I.K. Hong, Journal of Industrial and Engineering 2249.
Chemistry 16 (2010) 1006. [63] S. Puhan, N. Vedaraman, G. Sankaranarayanan, V. Boppan, B. Ram, Renewable
[5] S.B. Lee, J.D. Lee, I.K. Hong, Journal of Industrial and Engineering Chemistry 17 Energy 30 (2005) 1269.
(2011) 138. [64] S.V. Ghadge, H. Raheman, Biomass & Bioenergy 28 (2005) 601.
[6] S.S. Kim, K.H. Kim, S.C. Shin, E.S. Yim, Journal of the Korean Industrial and [65] V. Kumari, S. Shah, M.N. Gupta, Energy & Fuels 21 (2007) 368.
Engineering Chemistry 18 (5) (2007) 401. [66] J. Zhang, L. Jiang, Bioresource Technology 99 (2008) 8995.
[7] B. Saeid, M.K. Aroua, A. Abdul Raman, N.M.N. Sulaiman, Journal of Chemical and [67] M.J. Haas, P.J. Michalski, S. Runyon, A. Nunez, K.M. Scott, Journal of the American
Engineering Data 53 (2008) 877. Oil Chemists Society 80 (2003) 97.
[8] K.R. Szulczyk, B.A. McCarl, Renewable & Sustainable Energy Reviews 14 (2010) [68] L.C. Meher, V.S.S. Dharmagadda, S.N. Naik, Bioresource Technology 97 (2006)
2426. 1392.
[9] A.B. Chhetri, K.C. Watts, M.R. Islam, Energies 1 (2008) 3. [69] J. Van Gerpen, Fuel Processing Technology 86 (2005) 1097.
[10] M.J. Haas, M.S. Karen, W.N. Marmer, T.A. Foglia, Journal of the American Oil [70] K. Nie, F. Xie, F. Wang, T. Tianwei, Journal of Molecular Catalysis B: Enzymatic 43
Chemists Society 81 (2004) 83. (2006) 142.
[11] R. Bai, S. Wang, F. Mei, T. Li, G. Li, Journal of Industrial and Engineering Chemistry [71] W. Davies, Biodiesel Technologies and Plant Design, August 2005 davies@sn2.-
17 (2011) 777. com.au.
[12] J.M. Encinar, J.F. González, A. Rodrı́guez-Reinares, Fuel Processing Technology 88 [72] B.H. Uma, Y.S. Kim, Journal of Industrial and Engineering Chemistry 15 (2009) 1.
(2007) 513. [73] T. Minowa, S.Y. Yokoya, M. Kishimoto, T. Okakura, Fuel 74 (1995) 1731.
[13] Y.C. Sharma, B. Singh, S.N. Upadhyay, Fuel 87 (2008) 2355. [74] H. GuanHua, F. Chen, D. Wei, X.W. Zhang, G. Chen, Applied Energy 87 (2010) 38.
[14] F. Hideki, K. Akihiko, N. Hideo, Journal of Bioscience and Bioengineering 92 [75] M. Xin, J. Yang, X. Xu, L. Zhang, Q. Nie, M. Xian, Renewable Energy 34 (2009) 1.
(2001) 405. [76] S.A. Khan, Rashmi, M.Z. Hussain, S. Prasad, U.C. Banerjee, Renewable & Sustain-
[15] Z. Helwani, M.R. Othman, N. Aziz, W.J.N. Fernando, J. Kim, Fuel Processing able Energy Reviews 13 (2009) 2361.
Technology 90 (2009) 1502. [77] S.A. Scott, M.P. Davey, J.S. Dennis, I. Horst, C.J. Howe, D.J. Lea-Smith, A.G. Smith,
[16] A. Demirbas, Energy Conversion and Management 50 (2009) 14. Current Opinion in Biotechnology 21 (2010) 277.
[17] A. Demirbas, A Realistic Fuel for Alternative Diesel Engines, Springer, ISBN 978- [78] T.M. Mata, A.A. Martins, N.S. Caetano, Renewable & Sustainable Energy Reviews
1-84628-994-1, e-ISBN 978-1-84628-995-8, doi:10.1007/978-1-84628-995-8. 14 (2010) 217.
[18] J. Van Gerpen, B. Shanks, R. Pruszko, D. Clements, G. Knothe, Biodiesel Production [79] A. Demirbas, Energy Conversion and Management 51 (2010) 2738.
Technology, NREL/SR-510-36244 August 2002–January 2004. [80] J. Singh, S. Gu, Renewable & Sustainable Energy Reviews 14 (2010) 1367.
[19] S. Chongkhong, C. Tongurai, P. Chetpattananondh, Renewable Energy 34 (2009) [81] A. Demirbas, M.F. Demirbas, Energy Conversion and Management 52 (2011) 163.
1059. [82] A. Singh, P.S. Nigam, J.D. Murphy, Bioresource Technology 102 (2011) 10.
[20] K. Bozbas, Renewable & Sustainable Energy Reviews 12 (2008) 542.
26 I.M. Atadashi et al. / Journal of Industrial and Engineering Chemistry 19 (2013) 14–26

[83] N.H. Tran, J.R. Bartlett, G.S.K. Kannangara, A.S. Milev, H. Volk, M.A. Wilson, Fuel [137] M. Di Serio, M. Ledda, M. Cozzolino, G. Minutillo, R. Tesser, E. Santacesaria,
89 (2010) 265. Industrial and Engineering Chemistry Research 45 (9) (2006) 3009.
[84] S. Behzadi, M.M. Farid, Bioresource Technology 100 (2009) 683. [138] W. Xie, H. Peng, L. Chen, Journal of Molecular Catalysis A: Chemical 246 (2006)
[85] N.B. Haq, M.A. Hanif, M. Qasim, Ata-ur-Rehman, Fuel 87 (2008) 2961. 24.
[86] A. Casas, M.J. Ramos, A. Pérez, Biomass & Bioenergy (2011) 1. [139] Z. Wen, X. Yu, S.-T. Tu, J. Yan, E. Dahlquist, Applied Energy 87 (2010) 743.
[87] N.A. Khan, H. el Dessouky, Renewable & Sustainable Energy Reviews 13(2009) [140] O. Meyer, F. Rößner, R.A. Rakoczy, R.W. Fischer, DGMK Conf, September 29–
1576. October 1, Berlin, Germany, 2008.
[88] Y. Wang, S. Ou, P. Liu, Z. Zhang, Energy Conversion and Management 48 [141] J.R. Marı́a, A. Casas, L. Rodrı́guez, R. Romero, A. Pérez, Applied Catalysis A:
(2007) 184. General 346 (2008) 79.
[89] G. Vicente, M. Martinez, J. Aracil, Bioresource Technology 92 (2004) 297. [142] N. Supamathanon, J. Wittayakun, S. Prayoonpoka, Journal of Industrial and
[90] R. Umer, F. Anwara, B.R. Moser, S. Ashrafa, Biomass & Bioenergy 32 (2008) 1202. Engineering Chemistry 17 (2011) 182.
[91] A. Srivastava, R. Prasad, Renewable & Sustainable Energy Reviews 4 (2000) 111. [143] S. Praserthdam, P. Wongmaneenil, B. Jongsomjit, Journal of Industrial and
[92] A.K. Agarwal, Progress in Energy and Combustion Science 33 (2007) 233. Engineering Chemistry 16 (2010) 935.
[93] K.-H. Chung, Journal of Industrial and Engineering Chemistry 16 (2010) 506. [144] J. Kathlene, R. Gopinath, L.C. Meher, A.D. Kumar, Applied Catalysis B: Environ-
[94] D. Kusdiana, S. Saka, Bioresource Technology 91 (2004) 289. mental 85 (2008) 86.
[95] J.M. Marchetti, A.F. Errazu, Biomass & Bioenergy 32 (2008) 892. [145] H.L. Ngo, N.A. Zafiropoulos, T.A. Foglia, E.T. Samulski, W. Lin, Journal of the
[96] A.A. Refaat, S.T. El Sheltawy, K.U. Sadek, International Journal of Environmental American Oil Chemists Society 87 (2010) 445.
Science and Technology 5 (3) (2008) 315. [146] S. Furuta, H. Matsuhasbi, K. Arata, Catalysis Communications 5 (2004) 721.
[97] N. Saifuddin, K.H. Chua, Malaysian Journal of Chemistry 6 (2004) 077. [147] J. Jitputti, B. Kitiyanan, P. Rangsunvigit, K. Bunyakiat, L. Attanatho, P. Jenvanit-
[98] F. Ferella, G.M. Di Celso, I. De Michelis, V. Stanisci, F. Vegliò, Fuel 89 (2010) 36. panjakul, Chemical Engineering Journal 116 (2006) 61.
[99] L. Wang, H. He, Z. Xie, J. Yang, S. Zhu, Fuel Processing Technology 8 (2007) 8477. [148] P. Wongmaneenil, B. Jongsomjit, P. Praserthdam, Journal of Industrial and
[100] S. Induri, S. Sengupta, J.K. Basu, Journal of Industrial and Engineering Chemistry Engineering Chemistry 16 (2010) 327.
16 (2010) 467. [149] D.E. Lopez, J.G. Goodwin Jr., D.A. Bruce, Journal of Catalysis 245 (2007) 381.
[101] N.U. Soriano Jr., R. Venditti, D.S. Argyropoulos, Fuel 88 (2009) 560. [150] F. Chai, F. Cao, F. Zhai, Y. Chen, X. Wang, Z. Su, Advanced Synthesis and Catalysis
[102] A.L. Cardoso, S. Cristina, G. Neves, M.J. Da Silva, Energies 1 (2009) 79. 349 (2007) 1057.
[103] S. Gryglewicz, Bioresource Technology 70 (1999) 249. [151] Rohm and Hass, AMBERLYSTTM BD20 solid catalyst FFAs Esterification Technol-
[104] G. Antolın, F.V. Tinaut, Y. Briceno, V. Castano, C. Perez, A.I. Ramırez, Bioresource ogy Philadelphia, USA. www.amberlyst.com.
Technology 83 (2002) 111. [152] M. Di Serio, M. Cozzolino, R. Tesser, P. Patrono, F. Pinzari, B. Bonelli, E. Santa-
[105] L. Canoira, R. Alcántara, M.J. Garcı́a-Martı́nez, J. Carrasco, Biomass & Bioenergy cesaria, Applied Catalysis A: General 320 (2007) 1.
30 (2006) 76. [153] K. Chung, D. Chang, B. Park, Bioresource Technology 99 (2008) 7438.
[106] M. Cardone, M. Mazzoncini, S. Menini, V. Roccoc, A. Senatorea, M. Seggianid, S. [154] H.J. Park, J.-I. Dong, J.-K. Jeon, K.-S. Yoo, J.-H. Yim, J.M. Sohn, Y.-K. Park, Journal of
Vitolod, Biomass & Bioenergy 25 (2003) 623. Industrial and Engineering Chemistry 13 (2007) 182.
[107] S.L. Dmytryshy, A.K. Dalai, S.T. Chaudhari, H.K. Mishra, M.J. Reaney, Bioresource [155] S.J. Kim, M.J. Park, K.-D. Jung, O.-S. Joo, Journal of Industrial and Engineering
Technology 92 (2004) 55. Chemistry 10 (2004) 995.
[108] N. Foidl, G. Foidl, M. Sanchez, M. Mittelbach, S. Hackel, Bioresource Technology [156] K.-H. Chung, B.-G. Park, Journal of Industrial and Engineering Chemistry 15
58 (1996) 77. (2009) 388.
[109] M.N. Nabi, R. Md Mustafizur, A. Md Shamim, Applied Thermal Engineering 29 [157] A. Salis, M. Pinna, M. Monduzzi, V. Solinas, Journal of Molecular Catalysis B:
(2009) 2265. Enzymatic 54 (2008) 19.
[110] P. Nakpong, W. Sasiwimol, Fuel 89 (2010) 1806. [158] K. Nie, F. Xie, F. Wang, T. Tan, Journal of Molecular Catalysis B: Enzymatic 43
[111] O.E. Ikwuagwu, I.C. Ononogbu, O.U. Njoku, Industrial Crops and Products 12 (2006) 142.
(2000) 57. [159] C.J. Shieh, H.F. Liao, C.C. Lee, Bioresource Technology 88 (2003) 103.
[112] A.V. Tomasevic, S.S.S. Marinkovic, Fuel Processing Technology 81 (2003) 1. [160] A.V.L. Pizarro, E.Y. Park, Process Biochemistry (Oxford, U.K.) 38 (2003) 1077.
[113] N. Usta, Biomass & Bioenergy 28 (2005) 77. [161] C.A. Casimir, C. Shu-wei, L. Guan-chiun, J. Shaw, Journal of Agricultural and Food
[114] P.D. Prafulla, S. Deng, Fuel 88 (2009) 1302. Chemistry 55 (2007) 8995.
[115] L.C. Meher, M.G. Kulkarni, A.K. Dalai, S.N. Naik, European Journal of Lipid Science [162] F. Abreu, F.B. Alves, C.C.S. Macêdo, L.F. Zara, P.A.Z. Suarez, Journal of Molecular
and Technology 108 (2006) 5389. Catalysis A: Chemical 227 (2005) 263.
[116] L.Y. Wang, J.C. Yang, Fuel 86 (3) (2007) 328. [163] M.C.G. Albuquerque, I. Jimenez-Urbistondo, J. Santamarıa-Gonzalez, J.M. Mer-
[117] Chementator, Biodiesel Production Using a Heterogeneous Catalyst, January ida-Robles, R. Moreno-Tost, E. Rodrıguez-Castellon, A. Jimenez-Lopez, D.C.S.
2008, http://www.axens.net/upload/news/fichier/chemical_engineering.pdf, Azevedo, C.L. Cavalcante Jr., P. Maireles-Torres, Applied Catalysis A 334
2004. (2008) 35.
[118] K.G. Georgogianni, A.P. Katsoulidis, P.J. Pomonis, M.G. Kontominas, Fuel Proces- [164] M.J. Ramos, C. Abraham, R. Lourdes, R. Rubı́, P. Ángl, Applied Catalysis A: General
sing Technology 90 (2009) 671. 346 (2008) 79.
[119] J.G. Suppes, M.A. Dasari, E.J. Doskocil, P.J. Mankidy, M.J. Goff, Applied Catalysis A: [165] H.J. Kim, B.S. Kang, M.J. Kim, Y.M. Park, D.K. Kim, J.S. Lee, K.Y. Lee, Catalysis Today
General 257 (2004) 213. 93–5 (2004) 315.
[120] http://www.axens.net/html-gb/offer/offer_processes_104.html (assessed [166] K. Jacobson, G. Rajesh, C.M. Lekha, K.D. Ajay, Applied Catalysis B: Environmental
3/7/2011). 85 (2008) 86.
[121] S. Semwal, A.K. Arora, R.P. Badoni, D.K. Tuli, Bioresource Technology 102 (2011) 2151. [167] N. Boz, N. Degirmenbasi, D.M. Kalyon, Applied Catalysis B: Environmental 89
[122] J.-S. Lee, S. Saka, Bioresource Technology 101 (2010) 7191. (2009) 590.
[123] M. Kouzu, T. Kasuno, M. Tajika, Y. Sugimoto, S. Yamanaka, J. Hidaka, Fuel 87 [168] H. Noureddini, X. Gao, R.S. Philkana, Bioresource Technology 96 (2005) 769.
(2008) 2798. [169] G. Sunita, B.M. Devassy, A. Vinu, D.P. Sawant, V.V. Balasubramanian, S.B. Halli-
[124] W.M. Antunes, C.O. Veloso, C.A. Henriques, Catalysis Today 133–135 (2008) 548. gudi, Catalysis Communications 9 (2008) 696.
[125] X. Liu, H. He, Y. Wang, S. Zhu, Fuel 87 (2008) 216. [170] S. Saka, Y. Isayama, Fuel 88 (2009) 1307.
[126] B.P. Lim, G.P. Maniam, S. Abd Hamid, European Journal of Scientific Research [171] G. Madras, C.R. Kolluru, Fuel 83 (2004) 2029.
(2009) 347. [172] D. Martı́n Alonso, R. Mariscal, R. Moreno-Tost, M.D. Zafra Poves, M. López
[127] J.F. Puna, J.F. Gomes, M. Joana, N. Correia, A.P. Soares Dias, J.C. Bordado, Fuel 89 Granados, Catalysis Communications 8 (2007) 2074.
(2010) 3602. [173] J. Janaun, N. Ellis, Sustainable Energy Reviews 14 (2010) 1312.
[128] Y.C. Sharma, B. Singh, J. Korstad, Fuel 90 (2011) 1309. [174] B.H. Hameed, L.F. Lai, L.H. Chin, Fuel Processing Technology 90 (2009) 606.
[129] T.-M. Hsin, S. Chen, E. Guo, C.-H. Tsai, M. Pruski, V.S.Y. Lin, Topics in Catalysis 53 [175] M. Mittelbach, A. Silberholz, M. Koncar, in: Proceedings of the 21st World
(2010) 746. Congress of the International Society for Fats Research, The Hague, October,
[130] M. Zabeti, W.M.A.W. Daud, M.K. Aroua, Fuel Processing Technology 91 (2010) (1996), p. 497.
243. [176] D. Lee, Y. Park, K. Lee, Catalysis Surveys from Asia 13 (2009) 63.
[131] O. Ilgen, A.N. Akin, Turkish Journal of Chemistry 33 (2009) 1. [177] A. Demirbas, Energy Conversion and Management 47 (2006) 2271.
[132] K. Schulze, W. Makowski, R. Chyzy, R. Dziembaj, G. Geismar, Applied Clay [178] Z.J. Predojevic, Fuel 87 (2008) 3522.
Science 18 (2001) 59. [179] I.M. Atadashi, M.K. Aroua, A. Abdul Aziz, Renewable Energy 36 (2011) 437.
[133] J.L. Shumaker, C. Crofcheck, S.A. Tackett, E. Santillan-Jimenez, T. Morgan, Y. Ji, M. [180] D. Casanave, D. Jean-Luc, E. Freund, Pure and Applied Chemistry 79 (2007) 2071.
Crocker, T.J. Toops, Applied Catalysis B: Environmental 82 (2008) 120. [181] Z. Yang, W. Xie, Fuel Processing Technology (2007) 631.
[134] N. Barakos, S. Pasias, N. Papayannakos, Bioresource Technology 99 (2008) 5037. [182] H. Gerard, B. Delfort, D. le Pennec, L. Bournay, C. Jean-Alain, Preprints of Papers:
[135] H.-Y. Zeng, Z. Feng, X. Deng, Y.-Q. Li, Fuel 87 (13–14) (2008) 3071. American Chemical Society, Division of Fuel Chemistry 48 (2) (2003) 636.
[136] D. Siano, M. Nastasi, E. Santacesaria, M. Di Serio, R. Tesser, G. Minutillo, M. Ledda, [183] I.M. Atadashi, M.K. Aroua, A. Abdul Aziz, Renewable & Sustainable Energy
T. Tenore, WO 050925 A1 (2006). Reviews 14 (2010) 1999.

Anda mungkin juga menyukai