Você está na página 1de 13

Applied Energy 87 (2010) 10831095

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

A review on biodiesel production using catalyzed transesterication


Dennis Y.C. Leung *, Xuan Wu, M.K.H. Leung
Department of Mechanical Engineering, The University of Hong Kong, Pokfulam Road, Hong Kong, China

a r t i c l e

i n f o

Article history:
Received 17 June 2009
Received in revised form 6 October 2009
Accepted 7 October 2009
Available online 7 November 2009
Keywords:
Biodiesel
Alkali-catalyzed transesterication
Feedstock
Purication
Mass transfer

a b s t r a c t
Biodiesel is a low-emissions diesel substitute fuel made from renewable resources and waste lipid. The
most common way to produce biodiesel is through transesterication, especially alkali-catalyzed transesterication. When the raw materials (oils or fats) have a high percentage of free fatty acids or water, the
alkali catalyst will react with the free fatty acids to form soaps. The water can hydrolyze the triglycerides
into diglycerides and form more free fatty acids. Both of the above reactions are undesirable and reduce
the yield of the biodiesel product. In this situation, the acidic materials should be pre-treated to inhibit
the saponication reaction. This paper reviews the different approaches of reducing free fatty acids in the
raw oil and renement of crude biodiesel that are adopted in the industry. The main factors affecting the
yield of biodiesel, i.e. alcohol quantity, reaction time, reaction temperature and catalyst concentration,
are discussed. This paper also described other new processes of biodiesel production. For instance, the
Biox co-solvent process converts triglycerides to esters through the selection of inert co-solvents that
generates a one-phase oil-rich system. The non-catalytic supercritical methanol process is advantageous
in terms of shorter reaction time and lesser purication steps but requires high temperature and pressure. For the in situ biodiesel process, the oilseeds are treated directly with methanol in which the catalyst has been preciously dissolved at ambient temperatures and pressure to perform the
transesterication of oils in the oilseeds. This process, however, cannot handle waste cooking oils and
animal fats.
2009 Elsevier Ltd. All rights reserved.

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Biodiesel production with catalyzed transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Basic chemical reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Biodiesel production processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1.
Process flow chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2.
Raw materials treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.3.
Pretreatment of acidic feedstocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.4.
Catalyst and alcohol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.5.
Mixing and neutralization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.6.
Transesterification and separation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.7.
Refining crude glycerol. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.8.
Purification of crude biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.9.
Quality control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Storage of biodiesel product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Main factors affecting the yield of biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.1.
Alcohol quantity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.2.
Reaction time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.3.
Reaction temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.4.
Catalyst concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Other processes of biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Biox co-solvent process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

* Corresponding author. Tel.: +00 852 2859 7911; fax: +00 852 2858 5415.
E-mail address: ycleung@hku.hk (D.Y.C. Leung).
0306-2619/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apenergy.2009.10.006

1084
1085
1085
1086
1086
1086
1087
1087
1088
1088
1088
1089
1089
1090
1091
1091
1091
1091
1091
1091
1091

1084

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

4.

3.2.
Supercritical alcohol process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
In situ biodiesel process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction

1091
1091
1092
1092
1092

oils which are not suitable for human consumption because of the
presence of some toxic components in the oils. Furthermore, nonedible oil crops can be grown in waste lands that are not suitable
for food crops and the cost of cultivation is much lower because
these crops can still sustain reasonably high yield without intensive care [12,26]. However, most non-edible oils contain high free
fatty acids. Thus they may require multiple chemical steps or alternate approaches to produce biodiesel, which will increase the production cost, and may lower the ester yield of biodiesel below the
standards [14,25,27]. Animal fats contain higher saturated fatty
acids and normally exist in solid form at room temperature that
may cause problems in the production process. Its cost is also higher than vegetable oils [18]. UFO is not suitable for human consumption but is a feedstock for biodiesel production. Its usage
signicantly reduces the cost of biodiesel production. However,
the quality of UFO may cause concern because its physical and
chemical properties depend on the contents of fresh cooking oil
and UFO may contain lots of undesired impurity, such as water,
free fatty acids [18,26,28,29]. Since the cost of raw materials accounts about 6080% of the total cost of biodiesel production,
choosing a right feedstock is very important [18,26]. Also, the yield
and properties of biodiesel products produced from different feedstocks would be quite different from each other. Table 2 shows
some physicochemical properties of biodiesel from different feedstocks and their yields under different production conditions
[7,9,11,14,16,18,20,22,2528,3038]. Although at present biodiesel

Due to the depletion of the worlds petroleum reserves and the


increasing environmental concerns, there is a great demand for
alternative sources of petroleum-based fuel, including diesel and
gasoline fuels. Biodiesel, a clean renewable fuel, has recently been
considered as the best candidate for a diesel fuel substitution because it can be used in any compression ignition engine without
the need for modication [1]. Chemically, biodiesel is a mixture
of methyl esters with long-chain fatty acids and is typically made
from nontoxic, biological resources such as vegetable oils [224],
animal fats [16,17,23,24], or even used cooking oils (UFO) [10]. Table 1 shows some feedstocks of biodiesel and their physicochemical properties. Vegetable oils are promising feedstocks for biodiesel
production since they are renewable in nature, and can be produced on a large scale and environmentally friendly [25]. Vegetable oils include edible and non-edible oils. More than 95% of
biodiesel production feedstocks comes from edible oils since they
are mainly produced in many regions and the properties of biodiesel produced from these oils are much suitable to be used as diesel
fuel substitute [26]. However, it may cause some problems such as
the competition with the edible oil market, which increases both
the cost of edible oils and biodiesel [11]. Moreover, it will cause
deforestation in some countries because more and more forests
have been felled for plantation purposes. In order to overcome
these disadvantages, many researchers are interested in non-edible

Table 1
Feedstocks for biodiesel production and their physicochemical properties.
Type of
oil

Species

Vegetable oil
Edible oil Soybean
Rapessed
Sunower
Palm
Peanut
Corn
Camelina

Nonedible
oil

Canola
Cotton
Pumpkin
Jatropha
curcas
Pongamina
pinnata
Sea mango
Palanga
Tallow
Nile tilapia
Poultry

Others
a
b

Used
cooking oil

Main chemical composition (fatty


acid composition wt.%)

Density
(g/cm3)

Flash
point (C)

Kinematic viscosity
(cst, at 40 C)

Acid value
(mg KOH/g)

Heating value
(MJ/kg)

References

C16:0,
C16:0,
C16:0,
C16:0,
C16:0,
C22:0
C16:0,
C16:0,
C18:3,
C16:0,
C16:0,
C16:0,
C16:0,

0.91
0.91
0.92
0.92
0.90

254
246
274
267
271

32.9
35.1
32.6
39.6a
22.72

0.2
2.92

0.1
3

39.6
39.7
39.6

39.8

[6,15,18,19]
[3,8,15,18]
[15,18,20]
[2,11,18]
[9,13,18,20]

39.5
42.2

[17,18]
[5,21]
[8,10]
[17,18]
[22]
[4,7,12]

C18:1,
C18:0,
C18:0,
C18:0,
C18:0,

C18:2
C18:1,
C18:1,
C18:1,
C18:1,

C18:2
C18:2
C18:2
C18:2, C20:0,

C18:0,
C18:0,
C20:0,
C18:0,
C18:0,
C18:0,
C16:1,

C18:1,
C18:1,
C20:1,
C18:1,
C18:1,
C18:1,
C18:0,

C18:2, C18:3
C18:2
C20:3
C18:2, C18:3
C18:2
C18:2
C18:1, C18:2

0.91
0.91

277

34.9a

0.76

234
>230
225

38.2
18.2
35.6
29.4

0.4

0.91
0.92
0.92

0.55
28

39.5
39
38.5

C16:0, C18:0, C18:1, C18:2, C18:3

0.91

205

27.8

5.06

34

[14]

C16:0, C18:0, C18:1, C18:2


C16:0, C18:0, C18:1, C18:2
C14:0, C16:0, C16:1, C17:0, C18:0,
C18:1, C18:2
C16:0, C18:1, C20:5, C22:6, other
acids
C16:0, C16:1, C18:0, C18:1, C18:2,
C18:3
Depends on fresh cooking oil

0.92
0.90
0.92

221

29.6
72.0

0.24
44

40.86
39.25
40.05

[11]
[14]
[17,23]

0.91

32.1b

2.81

[16]

0.90

39.4

[23,24]

0.90

44.7

2.5

[10]

Kinematic viscosity at 38 C, mm2/s.


Kinematic viscosity at 37 C, mm2/s.

1085

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

cannot entirely replace petroleum-based diesel fuel, there are several distinct advantages of biodiesel over diesel fuel. Biodiesel has
higher combustion efciency, cetane number than diesel fuel [39].
It is biodegradable and more than 90% biodiesel can be biodegraded within 21 days [40,41]. Biodiesel has lower sulfur and aromatic
content than diesel fuel and that means it will not emit lots of toxic
gas [42,43]. Moreover, it reduces most exhaust emissions except
NOx, such as monoxide, unburned hydrocarbons, and particulate
matter [4448].
A number of methods are currently available and have been
adopted for the production of biodiesel fuel. There are four primary
ways to produce biodiesel (Table 3): direct use and blending of raw
oils [4953], micro-emulsions [54], thermal cracking [5560], and
transesterication [61]. The most commonly used method for converting oils to biodiesel is through the transesterication of animal
fats or vegetable oils, which forms the basis of the present paper.
The objective of this paper is to give an overview on the technological advancement of producing biodiesel through transesterication. The main factors affecting the yield of biodiesel and other
types of transesterication will also be discussed.

transesterication. The simplied form of its chemical reaction is


presented in equation
CH2-O-CO-R1

CH2-OH

(Catalyst)

CH-O-CO-R2

R-O-CO-R 1

+ 3ROH

R-O-CO-R 2

CH-OH

CH2-O-CO-R3
(Triglyceride)

CH2-OH
(Alcohol)

(Glycerol)

R-O-CO-R 3
(Mixture of fatty acid esters)

2. Biodiesel production with catalyzed transesterication

where R1, R2, R3 are long-chain hydrocarbons, sometimes called


fatty acid chains. Normally, there are ve main types of chains in
vegetable oils and animal oils: palmitic, stearic, oleic, linoleic, and
linolenic. When the triglyceride is converted stepwise to diglyceride, monoglyceride, and nally to glycerol, 1 mol of fatty ester is liberated at each step [61]. Usually, methanol is the preferred alcohol
for producing biodiesel because of its low cost.
Vegetable oils and fats may contain small amounts of water and
free fatty acids (FFA). For an alkali-catalyzed transesterication,
the alkali catalyst that is used will react with the FFA to form soap.
Eq. (2) shows the saponication reaction of the catalyst (sodium
hydroxide) and the FFA, forming soap and water.

2.1. Basic chemical reactions

R1 COOH

Common vegetable oils or animal fats are esters of saturated


and unsaturated monocarboxylic acids with the trihydric alcohol
glyceride. These esters are called triglycerides, which can react
with alcohol in the presence of a catalyst, a process known as

This reaction is undesirable because the soap lowers the yield of


the biodiesel and inhibits the separation of the esters from the
glycerol. In addition, it binds with the catalyst meaning that more
catalyst will be needed and hence the process will involve a higher

FFA

NaOH

! R1 COONa H2 O

sodium hydroxide

soap

water

Table 2
Physicochemical properties of biodiesel from different oil source and the yields under the production conditions.
Feedstock

1
a
b

Kinematic
viscosity (cst,
at 40 C)

Density Saponication Iodine Acid


Cetane Heating
(g/cm3) number
value value
number value
(mg KOH/
(MJ/kg)
g)

Soybean

4.08

0.885

201

138.7

0.15

Rapeseed

4.35.83

Sunower
Palm

4.9
4.42

200
207

Peanut
Corn

4.42
3.39

Camelina

6.127a

Canola

3.53

Cotton
Pumpkin
Jatropha
curcas
Pongamina
pinnata
Sea mango

4.07
4.41
4.78

0.88
0.888
0.88
0.86
0.9
0.883
0.88
0.89
0.882
0.888
0.88
0.9
0.875
0.8837
0.8636

4.8

Production conditions1
T (C) P
M
(min)

C (wt.%)

40

65

90

1:12

0.250.45 4950

45

65

120

1:6

CaO|8 (2.03% H2O in


methanol)
KOH|1

142.7
60.07

0.24
0.08

49
62

45.3
34

60
Room

120
584

1:6
1:11

NaOH|1
KF/ZnO|5.52

200
202

67.45
120.3

54
5859

40.1
45

60
80

120
60

1:6
1:9

NaOH|0.5
KOH|2

0.080.52

Room

60

1:6

KOH|1.5

182

152
157
103.8

56

45

60

60

1:9

KOH|1

204
202
202

104.7
115
108.4

0.16
0.48
0.496

54

6163

45
38
4042

65
65
60

90
60
120

1:6
1:6
1:06

NaOH|0.75
NaOH|1
NaOH|1

0.883

0.62

6061

42

65

180

1:6

KOH|1

180

180

1:8

52

Palanga

3.99

0.869

41

66

1:12

Tallow
Nile tilapia
Poultry
Used
cooking
oil

0.856

0.867

244.5

251.23

126
88.1
130

0.65
1.4
0.25
0.15

59
51
61

60
30
50
60

1440
90
1440
20

1:30
1:9
1:30
1:7

Sulfated zirconiz
alumina|6
C6H5CH3 + H3PO4|5 + 5,
H2SO4|6.5; KOH|9
(three step)
H2SO4|2.5
KOH|2 (ultrasound)
H2SO4|1.25
NaOH|1.1

Yield
(%)

References

>95

[18,20,27,33,35]

95
[36]
96
97.1 [20,38]
89.23 [18,20,32]
89
85
96
97.9

[9,18,20]
[20,25]
[31]

80
95
96.9
97.5
98

[20,25]

97
98
83.8

[14,2526,34]

85

[14]

98.28
98.2
99.72
94.6

[30]
[16]
[30]
[28]

T = reaction temperature (C); P = transesterication reaction period (min); M = molecular ratio of oil/methanol; C = amount of catalyst (wt.%).
Kinematic viscosity at 20 C.
Volumetric ratio of oil/methanol.

[20,37]
[20,22]
[7,20,25]

[11]

1086

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

Table 3
Different methods of biodiesel production.
Methods

Denition

Advantage

Disadvantage

Problems of using in
engines

References

Direct use and


blending

Direct use as diesel fuel or blend with


diesel fuel

Liquid nature-portability

Higher viscosity

[913]

Heat content (80% of diesel fuel)


Readily available; renewability

Lower volatility
Reactivity of
unsaturated
hydrocarbon chains
Lower cetane number

Coking and trumpet


formation
Carbon deposits
Oil ring sticking;
thickening and gelling of
the lubricating oil
Irregular injector needle
sticking; incomplete
combustion
Heavy carbon deposits;
increase lubrication oil
viscosity

Micro-emulsions

Thermal cracking
(pyrolysis)
Transesterication

A colloidal equilibrium dispersion of


optically isotropic uid
microstructures with dimensions
generally in the 1150 nm range
formed spontaneously from two
immiscible liquids and one or more
ionic or non-ionic amphiphiles
The conversion of long-chain and
saturated substance (biomass basis)
to biodiesel by means of heat
The reaction of a fat or oil with an
alcohol in the presence of catalyst to
form esters and glycerol

Better spray patterns during


combustion
Lower fuel viscosities

Lower energy content

Chemically similar to petroleumderived gasoline and diesel fuel

Energy intensive and


hence higher cost

[1520]

Renewability; higher cetane


number; lower emissions; higher
combustion efciency

Disposal of byproduct (glycerol and


waste water)

[2123]

cost [62]. Water, originated either from the oils and fats or formed
during the saponication reaction, retards the transesterication
reaction through the hydrolysis reaction. It can hydrolyze the triglycerides to diglycerides and forms more FFA. The typical hydrolysis reaction is shown in Eq. (3).
CH2-O-CO-R1

CH2-OH

CH-O-CO-R2

+ H2O

CH-O-CO-R 2

R1-COOH

CH2-O-CO-R 3
(Triglyceride)

CH2-O-CO-R3
(Water)

[14,21]

(Diglyceride)

glycerides, free fatty acids, water, and other contaminants in various proportions. Some crude vegetable oils contain phospholipids
that need to be removed in a degumming step. Phospholipids can
produce lecithin, a commercial emulsier [65]. Liu et al. compared
the different degumming methods, such as membrane ltration,
hydration, acid micelles degumming, supercritical extraction, etc.
The characteristics of the raw oils should be investigated before
choosing the suitable degumming method because different
degumming methods have their advantages and disadvantages
[66]. The oil can be separated through membrane ltration according to the average molecular weight or the particle size of phospholipids. Although degumming method can solve the problem,

(FFA)

R1 COOH ROH ! ROCOR1 H2 O


FFA

alcohol

fatty acid ester

Oils or Fats
>

However, the FFA can react with alcohol to form ester (biodiesel) by an acid-catalyzed esterication reaction. This reaction is
very useful for handling oils or fats with high FFA, as shown in
the equation below:

FFA _ 2.5wt%

water

Normally, the catalyst for this reaction is concentrated sulphuric acid. Due to the slow reaction rate and the high methanol to oil
molar ratio that is required, acid-catalyzed esterication has not
gained as much attention as the alkali-catalyzed transesterication
[63].

FFA>2.5wt%

Catalyst
Pretreatment

Mixing
Alcohol

Neutralized oil

Transesterification

2.2. Biodiesel production processes

Phase separation
2.2.1. Process ow chart
Today, most of the biodiesel is produced by the alkali-catalyzed
process. Fig. 1 shows a simplied ow chart of the alkali-catalyst
process. As described earlier, feedstocks with high free fatty acid
will react undesirably with the alkali catalyst thereby forming
soap. The maximum amount of free fatty acids acceptable in an alkali-catalyzed system is below 2.5 wt.% FFA. If the oil or fat feedstock has a FFA content over 2.5 wt.%, a pretreatment step is
necessary before the transesterication process [64] (see Section
2.2.3).
2.2.2. Raw materials treatment
The raw materials, which can be vegetable oils, animal fats, or
recycled greases, used in the production of biodiesel contain tri-

Re-neutralization
Crude Biodiesel

Alcohol Recovery

Crude Glycerol

Purification

Quality Control

Biodiesel
Fig. 1. Simplied process ow chart of alkali-catalyzed biodiesel production.

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

its process involves two steps with the use of an organic solvent.
Therefore, this method has no superiority on cost because of its
complicated processing [67,68]. In the hydration process, because
of phospholipids hydrophilicity, hot water can be added into the
oil with stirring. The phospholipids solubility will be signicantly
reduced, so it could be separated from the oil by natural settlement. The hydration method features a simple process, easy operation, and high yields rening, but non-hydratable phospholipids
cannot be removed by this method, as reported by Verleyen et al.
[69]. For removing non-hydratable phospholipids, critic acid or
phosphoric acid can be added into the oil which is heated to
70 C, this is called the special micelles degumming method. Pan
et al. [70] found that the reaction time for this method is about
5 min, followed by neutralization through dilute lye. The supercritical extraction method is employed to separate the phospholipids.
Moreover, by rening with supercritical CO2 extraction, it can
effectively remove the free fatty acids and the peroxidation products that are in the crude oil. However, the high pressure required
in this process will be the biggest challenge in its industrial application [71,72].
The free fatty acids are removed in a rening step and excess
free fatty acids can be removed as soaps in a later pretreatment
step. In addition, deodorization is another important step in the
raw material treatment. During this step, steam, at 16 mm Hg
pressure, is injected into the oil at 490550 K in order to eliminate
free fatty acids, aldehydes, unsaturated hydrocarbons, and ketones,
all of which cause undesirable odors and avors in the oil [73].
Next, in order to determine the percentage of FFA in the oils or fats,
titration is performed. As described above, if the percentage of FFA
is over 2.5 wt.%, pretreatment is necessary to reduce the content of
FFA. This step also determines the amount of caustic soda required
in the neutralization step.
2.2.3. Pretreatment of acidic feedstocks
Many pretreatment methods have been proposed for reducing
the high free fatty acid content of the oils, including steam distillation [74], extraction by alcohol [75], and esterication by acidcatalysis [76]. However, steam distillation for reducing high free
fatty acids requires a high temperature and has low efciency. Because of the limited solubility of free fatty acids in alcohol, extraction by alcohol method needs a large amount of solvent and the
process is complicated. Compared with the two former methods,
esterication by acid-catalysis makes the best use of the free fatty
acids in the oil and transforms it into biodiesel [75,77]. The common pretreatment is esterication of the FFA with methanol in
the presence of acidic catalysts (usually sulphuric acid). The catalysts can be homogeneous acid-catalysts or solid acid-catalysts
[78]. Compared with the former one, solid acid-catalysts offer
some advantages for eliminating separation, corrosion, toxicity,
and environmental problems, but the reaction rate is slower [76].
As described earlier, free fatty acids will be converted to biodiesel
by direct acid esterication and the water needs to be removed. If
the acid value of the oils or fats is very high, one-step esterication
pretreatment may not reduce the FFA efciently because of the
high content of water produced during the reaction. In this case,
a mixture of alcohol and sulphuric acid can be added into the oils
or fats three times (three-step pre-esterication). The time required for this process is about 2 h and water must be removed
by a separation funnel before adding the mixture into the oils or
fats for esterication again [79]. Moreover, some researchers reduce the percent of FFA by using acidic ion exchange resins in a
packed bed. Strong commercial acidic ion exchange resins can be
used for the esterication of FFA in waste cooking oils but the loss
of the catalytic activity maybe a problem [8084].
An alternative approach to reduce the FFA is to use iodine as a
catalyst to convert free fatty acids into biodiesel. An obvious

1087

advantage of this approach is that the catalyst (iodine) can be recycled after the esterication reaction. Li et al. [85] found through
orthogonal tests, under the optimal conditions (i.e. iodine amount:
1.3 wt.% of oils; reaction temperature: 80 C; ratio of methanol to
oils: 1.75:1; reaction time: 3 h) that the FFA content can be reduced to <2% [85].
Another new method of pretreatment is to add glycerol into the
acidic feedstock and heat it to a high temperature of about 200 C,
normally with a catalyst such as zinc chloride. The glycerol will react with the FFA to form monoglycerides and diglycerides. Then
the FFA level will become low and biodiesel can be produced using
the traditional alkali-catalyzed transesterication method. The
advantage of this approach is that no alcohol is needed during
the pretreatment and the water formed from the reaction can be
immediately vaporized and vented from the mixture. However,
the drawbacks of this method are its high temperature requirement and relatively slow reaction rate [86].
2.2.4. Catalyst and alcohol
In general, there are three categories of catalysts used for biodiesel production: alkalis, acids, and enzymes [8789]. Enzyme
catalysts have become more attractive recently since it can avoid
soap formation and the purication process is simple to accomplish. However, they are less often used commercially because of
the longer reaction times and higher cost. To reduce the cost, some
researchers developed new biocatalysts in recent years. An example is so called whole cell biocatalysts which are immobilized
within biomass support particles. An advantage is that no purication is necessary for using these biocatalysts [9093]. Compare
with enzyme catalysts, the alkali and acid catalysts are more commonly used in biodiesel production [61]. The alkali and acid catalysts include homogeneous and heterogeneous catalysts. Due to
the low cost of raw materials, sodium hydroxide and potassium
hydroxide are usually used as alkali homogeneous catalysts and alkali-catalyzed transesterication is most commonly used commercially [36,38,9499]. These materials are the most economic
because the alkali-catalyzed transesterication process is carried
out under a low temperature and pressure environment, and the
conversion rate is high with no intermediate steps. However, the
alkali homogeneous catalysts are highly hygroscopic and absorb
water from air during storage. They also form water when dissolved in the alcohol reactant and affect the yield [28]. Therefore,
they should be properly handled. On the other hand, some heterogeneous catalysts are solid and it could be rapidly separated from
the product by ltration, which reduces the washing requirement.
In addition, solid heterogeneous catalysts can stimulatingly catalyze the transesterication and esterication reaction that can
avoid the pre-esterication step, thus these catalysts are particularly useful for those feedstocks with high free fatty acid content
[100]. However, using a solid catalyst, the reaction proceeds at a
slower rate because the reaction mixture constitutes a three-phase
system, which, due to diffusion reasons, inhibits the reaction
[33,101]. Table 4 classies the three categories of catalysts with
their advantages and disadvantages [27,33,82,102109].
The alcohol materials that can be used in the transesterication
process include methanol, ethanol, propanol, butanol, and amyl
alcohol. Among these alcohols, methanol and ethanol are used
most frequently. Methanol is especially used because of its lower
cost and its physical and chemical advantages. Ma and Hanna
[61] reported that methanol can react with triglycerides quickly
and the alkali catalyst is easily dissolved in it. However, due to
its low boiling point, there is a large explosion risk associated with
methanol vapors which are colorless and odorless. Both methanol
and methoxide are extremely hazardous materials that should be
handled carefully. It should be ensured that one is not exposed
to these chemicals during biodiesel production.

1088

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

Table 4
Advantages and disadvantages at different types of catalysts used in the biodiesel production.
Type

Example

Advantages

Disadvantages

References

Alkali
Homogeneous

NaOH, KOH

High catalytic activity, low cost,


favorable kinetics, modest
operation conditions
Noncorrosive, environmentally
benign, recyclable, fewer disposal
problems, easily separation, higher
selectivity, longer catalyst lifetimes

Low FFA requirement, anhydrous conditions,


saponication, emulsion formation, more
wastewater from purication, disposable
Low FFA requirement, anhydrous conditions,
more wastewater from purication, high
molar ratio of alcohol to oil requirement,
high reaction temperature and pressure,
diffusion limitations, high cost

[97,101104]

Concentrated sulphuric acid

Catalyze esterication and


transesterication simultaneously,
avoid soap formation

[97,102,103]

2
ZnO/I2, ZrO2 =SO2
4 , TiO2 =SO4 , carbon-based
solid acid catalyst, carbohydrate-derived
catalyst, Vanadyl phosphate, niobic acid,
sulphated zirconia, Amberlyst-15, NaonNR50
Candida antarctica fraction B lipase,
Rhizomucor mieher lipase

Catalyze esterication and


transesterication simultaneously,
recyclable, eco-friendly

Equipment corrosion, more waste from


neutralization, difcult to recycle, higher
reaction temperature, long reaction times,
weak catalytic activity
Low acid site concentrations, low
microporosity, diffusion limitations, high
cost

Expensive, denaturation

[27,82,104]

Heterogeneous

Acid
Homogeneous

Heterogeneous

Enzymes

CaO, CaTiO3, CaZrO3, CaOCeO2, CaMnO3,


Ca2Fe2O5, KOH/Al2O3, KOH/NaY, Al2O3/KI,
ETS-10 zeolite, alumina/silica supported
K2CO3

Avoid soap formation, nonpolluting,


easier purication

2.2.5. Mixing and neutralization


The purpose of mixing methanol with the catalyst is to produce
methoxide which reacts with the base oils. Most of the catalysts
(e.g. NaOH, KOH) are in solid form and do not readily dissolve into
methanol, it is best to start agitating the methanol in a mixer and
add the catalyst slowly and carefully [64]. Once the catalyst completely dissolves in the methanol, the methoxide is ready to be
added to the oil. Once the methoxide is added into the oil, a neutralization reaction will immediately start. Some alkali catalysts
will react with rudimental acids during the pretreatment step or
will react with the free fatty acids from the oil. Therefore, more catalyst needs to be added to complete the reaction.
2.2.6. Transesterication and separation
When the catalyst, alcohol, and oil are mixed and agitated in a
reaction vessel, a transesterication reaction will start. A stirred
reactor is usually used as the reaction vessel for continuous alkali-catalyzed biodiesel production. Recently, there is an increased
interest in new technologies related to mass transfer enhancement.
Maeda et al. [110] and Thanh et al. [111] produced biodiesel from
vegetable oils assisted by ultrasound which is a useful tool for
strengthening the mass transfer of immiscible liquids. Ultrasonic
irradiation causes cavitation of bubbles near the phase boundary
between immiscible liquid phases. The asymmetric collapse of
the cavitation bubbles disrupts the phase boundary and starts
emulsication instantly. Micro jets, formed by impinging one liquid to another, lead to intensive mixing of the system near the
phase boundary. With the use of ultrasound biodiesel can be produced without heating because the cavitation may lead to a localized increase in temperature at the phase boundary and enhance
the reaction [110112]. Moreover, Wen et al. [113] fabricated a
new reaction vessel Zigzag micro-channel reactor in recent years
and found that less energy consumption for biodiesel synthesis can
be achieved by using this reactor.
Leung and Guo [28] studied the effect of operating conditions
on the product yield and pointed out that heating the oil prior to
the mixing can increase the reaction rate and hence shorten the
reaction time. During this step, in order to speed up the reaction,
mixing brings the oil, the catalyst, and the alcohol into intimate
contact while the temperature is kept just below the boiling point
of the alcohol (i.e. 64.5 C for methanol). Normally, the reaction
pressure is close to the atmospheric pressure to prevent the loss

[33,99
101,103,104]

[82,98,103,104]

of alcohol, and excess alcohol is used to ensure total conversion


of the oil to its esters. As previously mentioned, if the free fatty acid
level or water level is too high, it may cause problems downstream
with the saponication and the separation of the glycerol by-product. Therefore, the amount of water and free fatty acids in the feedstock oil should be monitored during the reaction.
Once the transesterication reaction is completed, two major
products exist: esters (biodiesel) and glycerol. The glycerol phase
is much denser than the biodiesel phase and settles at the bottom
of the reaction vessel, allowing it to be separated from the biodiesel phase. Phase separation can be observed within 10 min and can
be completed within several hours of settling. The reaction mixture
is allowed to settle in the reaction vessel in order to allow the initial separation of biodiesel and glycerol, or the mixture is pumped
into a settling vessel. In some cases, a centrifuge may be used to
separate the two phases [86].
Both the biodiesel and glycerol are contaminated with an unreacted catalyst, alcohol, and oil during the transesterication step.
Soap that may be generated during the process also contaminates
the biodiesel and glycerol phase. Schumacher [114] suggested that
although the glycerol phase tends to contain a higher percentage of
contaminants than the biodiesel, a signicant amount of contaminants is also present in the biodiesel. Therefore, crude biodiesel
needs to be puried before use.
2.2.7. Rening crude glycerol
Although biodiesel is the desired product from the reactions,
the rening of glycerol is also important due to its numerous applications in different industrial products such as moisturizers, soaps,
cosmetics, medicines, and other glycerol products [115117]. It is
one of the few products that has a good reactivity on sump oil,
and is extremely effective for washing shearing shed oor, so it
can be used as a heavy duty detergent and degreaser. Whittington
[118] reported that glycerol can even be fermented to produce ethanol, which means more biofuel can be produced.
According to the statements of Van Gerpen et al. [86], typically
produced glycerol is about 50% glycerol or less in composition and
mainly contains water, salts, unreacted alcohol, and unused catalyst. The unused alkali catalyst is usually neutralized by an acid.
In some cases, hydrochloric or sulphuric acids are added into the
glycerol phase during the re-neutralization step and produce salts
such as sodium chloride or potassium sulphate, the latter can be

1089

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

recovered for use as a fertilizer [119]. Generally, water and alcohol


are removed to produce 8088% pure glycerol that can be sold as
crude glycerol. In more sophisticated operations, the glycerol is
distilled to 99% or higher purity and sold in different markets [120].
After the re-neutralization step, the alcohol in the glycerol phase
can be removed through a vacuum ash process or by other types of
evaporators. Usually, the alcohol vapor is condensed back into liquid and reused in the process. However, the alcohol may contain
water that should be removed in a distillation column before the
alcohol is returned to the process. The alcohol recovery step is more
difcult when the alcohol that is used, such as ethanol or isopropanol, forms an azeotrope with the water. Gerpen [121] proposed the
use of a molecular sieve to remove the water generated.
2.2.8. Purication of crude biodiesel
After separation from the glycerol phase, crude biodiesel is
mainly contaminated with residual catalyst, water, unreacted alcohol, free glycerol, and soaps that were generated during the transesterication reaction [114]. Normally, crude biodiesel enters a
neutralization step and then passes through an alcohol stripper before the washing step. In some cases, acid is added to crude biodiesel
to neutralize any remaining catalyst and to split any soap. Soaps react with the acid to form water soluble salts and free fatty acids. Gerpen [121] stated that neutralization before the washing step reduces
the materials required for the washing step and minimizes the potential for emulsions being formed during the washing step. Unreacted alcohol should be removed with distillation equipment
before the washing step to prevent excess alcohol from entering
the wastewater efuent [86]. The primary purpose of this step is to
wash out the remnants of the catalyst, soaps, salts, residual alcohol,
and free glycerol from the crude biodiesel. Generally, three main approaches are adopted for purifying biodiesel: water washing, dry
washing [122], and membrane extraction [123,124]. These approaches are briey shown in Table 5 and discussed in detail as
follows.
2.2.8.1. Water washing. Since both glycerol and alcohol are highly
soluble in water, water washing is very effective for removing both
contaminants. It also can remove any residual sodium salts and
soaps. The primary material for water washing is distilled warm
water or softened water (slightly acidic) [72,73]. Warm water prevents the precipitation of saturated fatty acid esters and retards the
formation of emulsions with the use of a gentle washing action.
Softened water (slightly acidic) eliminates calcium and magnesium

contamination and neutralizes any remaining alkali catalysts [86].


After washing several times, the water phase becomes clear, meaning that the contaminants have been completely removed. Then,
the biodiesel and water phases are separated by a separation funnel or centrifuge [125]. Moreover, because of the immiscibility of
water and biodiesel, molecular sieves and silica gels, etc., can also
be used to remove water from the biodiesel [86]. The remaining
water can be removed from the biodiesel by passing the product
over heated Na2SO4 (25 wt.% of the amount of the ester product)
overnight and then be removed by ltration [126]. However, there
are many disadvantages to this method, including an increased
cost and production time, polluting liquid efuent, product loss,
etc. Moreover, emulsions can form when washing the biodiesel
made from waste cooking oils or acidic feedstocks because of the
soap formation [127].
2.2.8.2. Dry washing. Cooke et al. [122] used dry washing by replacing the water with an ion exchange resin or a magnesium silicate
powder in order to remove impurities. These two dry washing
methods can bring the free glycerol level down and is reasonably
effective for removing soaps. Both the ion exchange process and
the magnesol process have the advantage of being waterless and
thus eliminate many of the problems outlined above. Although
the magnesol process has a better effect on the removal of methanol than the ion resins, none of the products from this process fulll the limits specied in the EN Standard [128,129].
2.2.8.3. Membrane extraction. Gabelman and Hwang [123] proved
that the contaminants can be removed by using a hollow ber
membrane extraction, such as polysulfone. In this method, a hollow ber membrane (1 m long, 1 mm diameter) lled with distilled
water is immersed into the reactor (20 C). The crude biodiesel is
pumped into the hollow ber membrane (ow rate: 0.5 ml/min;
operating pressure: 0.1 MPa). Following this step, biodiesel is
passed over heated Na2SO4 and then ltered to remove any
remaining water [124]. This approach effectively avoids emulsication during the washing step and decreases the loss during the
rening process. The purity of the biodiesel obtained is about
90% and the other properties conform to the ASTM standards
[130]. It is a very promising method for purifying biodiesel.
2.2.9. Quality control
For commercial fuel, the nished biodiesel must be analyzed
using sophisticated analytical equipment to ensure it meets inter-

Table 5
Different approaches for purifying crude biodiesel.
Approaches

Primary
material
used

Function

Phases separation

Advantage

Disadvantage

References

Water washing

Distilled
warm
water

Prevents precipitation
of saturated fatty acid
esters
Retards the emulsion
formation
Eliminates calcium
and magnesium
contamination
Brings the free
glycerol level down
and removing soaps

Separation funnel,
centrifuge, molecular
sieves, silica gels, etc.

Very effective in
removing contaminants

Increased cost and production


time, liquid efuent, product loss,
emulsions formation

[46,7274]

Waterless

Overruns the limit in the EN


Standard

[69,75,76]

Remove the
contaminants

Avoids the emulsion


formation and decreases
the rening loss

Probably high cost and low


throughput due to contaminants
existed

[70,71]

Softened
water
Dry washing

Membrane
extraction

Ion
exchange
resin
Magnesium
silicate
powder
Polysulfone

1090

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

national standards. A few specications have been set but the


ASTM D 6751 and EN 14214 standards are the most commonly
used standards. Even in blends with conventional diesel fuel, Mittelbach [131] stated that most people in the industry expect the
biodiesel blending stock to meet the relevant standard before
being blended. Some properties in the standard, such as the cetane
number or density, can reect the properties of the chemical compounds that make up the biodiesel, and other properties provide an
indication of the quality of the production process [54]. Generally,
biodiesel standards identify the parameters that pure biodiesel
must meet before being used as a pure fuel or being blended with
distillate fuels [120,129,130].
To ensure safe operation in diesel engines, the most important
aspects of the biodiesel product are the completion of the reaction,
the removal of the free glycerol, residual catalyst and alcohol, and
the absence of free fatty acids [39,41,42,132]. As mentioned before,
if the transesterication reaction is not complete then triglycerides, diglycerides, or monoglycerides may be left in the nal product. Chemically, each of these compounds contains a glycerol
molecule. Fuel with excessive free glycerol may plug the fuel lters
and cause combustion problems in the diesel engine. Therefore, the
ASTM standard requires the total glycerol to be <0.24% of the nal
biodiesel product [54,130]. On the other hand, since residual methanol, even as little as 1%, can lower the ashpoint of the nal biodiesel product from 170 C to <40 C, the EN 14214 standard limits
the amount of alcohol to a very low level [86,129]. Finally,
although a specic value for the residual catalyst is not included
in the ASTM standard, it is limited by the specication on levels
of sulfated ash, which may lead to engine deposits and high abrasive wear levels [42,130].
Because the European specication for sulfur content (i.e. EN
14214) is much tighter than the US requirement, Ali et al. [133] reported that a number of producers in Europe are resorting to the

use of vacuum distillation for the removal of sulfur compounds


from the nal biodiesel product. In addition, some vegetable oils,
yellow greases, and brown greases leave an objectionable color
in the biodiesel. Although there is no color specication in the
ASTM standard, in some cases, an activated carbon bed, which is
an effective method for the removal of excessive color, is used to
produce a colorless biodiesel [120].
2.3. Storage of biodiesel product
Biodiesel is safe to store and the properties of biodiesel should
conform to respective standards after it has been stored for a long
time. Table 6 shows the ATSM D 6751 and EN 14214 standards.
There are several key factors that need to be considered for the
storage of biodiesel, including exposure temperature, oxidative
stability, fuel solvency, and material compatibility [40,134]. Lee
et al. [135] stated that the temperature of stored biodiesel should
be controlled so as to avoid the formation of crystals which can
plug fuel lines and fuel lters. For this reason, the storage temperature of most pure biodiesel is generally kept between 7 and 10 C.
Even in extremely cold climates, underground storage of pure biodiesel usually provides the storage temperature necessary for preventing crystal formation [86].
Bondioli et al. [136] noted that the stability of biodiesel is an
important property when it is to be stored for a prolonged period.
Poor stability can lead to an increased acid value and fuel viscosity
and to the formation of gums and sediments. Therefore, if the duration of storing biodiesel or biodiesel blends is more than 6 months,
it should be treated with an antioxidant additive [86]. Moreover,
because water contamination will lead to biological growth in
the fuel, it should be minimized in the stored fuel by using biocides. Biodiesel storage tanks made of aluminum, steel, Teon,
and uorinated polyethylene or polypropylene should be selected.

Table 6
Specications and test methods of ASTM D6751 and EN 14214 standards [129,130].
Property

Flash point
Kinematic viscosity at 40 C
Cetane number
Sulphated ash content
Copper strip corrosion
Acid value
Free glycerol
Total glycerol
Phosphorous content
Carbon residue
ASTM D6751 (100% sample)
EN 14214 (10% bottoms)
Cloud point
Density at 15 C
Distillation T90 AET
Sulfur (S 15 Grade)
Sulfur (S 500 Grade)
Sulfur content
Water and sediment
Water content
Total contamination
Oxidation stability at 110 C
Iodine value
Linolenic acid methyl ester
Polyunsaturated (P4 double bonds) methyl esters
Ester content
Methanol content
Monoglyceride content
Diglyceride content
Triglyceride content
Alkaline metals (Na + K)

Unit

Limits

Test method

ASTM D6751

EN 14214

ASTM D6751

EN 14214

C
mm2/s

% (m/m)

mg KOH/g
% (m/m)
% (m/m)
% (m/m)

130.0 min
1.96.0
47 min
0.020 max
No. 3 max
0.80 max
0.020 max
0.240 max
0.001 max

101.0 min
3.55.0
51 min

0.25 max
0.01 max

D93
D445
D613
D874
D130
D664
D6584
D6584
D4951

ISO CD3679e
EN ISO 3104
EN ISO 5165
ISO 3987
EN ISO 2160
pr EN 14104
pr EN 14105m pr EN 14106
pr EN 14105m
pr EN 141101

% (m/m)

0.050 max

Report customer

360 max
0.0015 max
0.05 max

0.050 max

0.3 max

860900

10 max

500 max
24 max
6 min
120 max
12 max
1 max
96.5 min
0.2 max
0.8 max
0.2 max
0.2 max
5 max

D4530

D2500

D1160
D5453
D5453

D2709

EN ISO 10370

EN SIO 3675 EN SIO 12185

EN ISO 12937
EN 12662
pr EN 14112
pr EN 14111
pr EN 14103d
pr EN 14103
pr EN 14103d
pr EN 141101
pr EN 14105m
pr EN 14105m
pr EN 14105m
pr EN 14108 pr EN 14109

C
kg/m3
C
ppm
ppm
mg/kg
%vol.
mg/kg
mg/kg
h

% (m/m)
% (m/m)
% (m/m)
% (m/m)
% (m/m)
% (m/m)
% (m/m)
mg/kg

Class 1
0.5 max

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

The tanks should minimize the possibility of water contamination


and should be cleaned prior to use for biodiesel storage [137].
2.4. Main factors affecting the yield of biodiesel
2.4.1. Alcohol quantity
Many researchers recognized that one of the main factors
affecting the yield of biodiesel is the molar ratio of alcohol to triglyceride [28,47,61,138]. Theoretically, the ratio for transesterication reaction requires 3 mol of alcohol for 1 mol of triglyceride to
produce 3 mol of fatty acid ester and 1 mol of glycerol. An excess
of alcohol is used in biodiesel production to ensure that the oils
or fats will be completely converted to esters and a higher alcohol
triglyceride ratio can result in a greater ester conversion in a shorter time. The yield of biodiesel is increased when the alcohol triglyceride ratio is raised beyond 3 and reaches a maximum. Further
increasing the alcohol amount beyond the optimal ratio will not increase the yield but will increase cost for alcohol recovery [28]. In
addition, the molar ratio is associated with the type of catalyst
used and the molar ratio of alcohol to triglycerides in most investigations is 6:1, with the use of an alkali catalyst [47,138]. When
the percentage of free fatty acids in the oils or fats is high, such
as in the case of waste cooking oil, a molar ratio as high as 15:1
is needed when using acid-catalyzed transesterication
[28,133,139,140].
2.4.2. Reaction time
Freedman et al. [141] found that the conversion rate of fatty
acid esters increases with reaction time. At the beginning, the reaction is slow due to the mixing and dispersion of alcohol into the oil.
After a while, the reaction proceeds very fast. Normally, the yield
reaches a maximum at a reaction time of <90 min, and then remains relatively constant with a further increase in the reaction
time [28,142]. Moreover, excess reaction time will lead to a reduction in the product yield due to the backward reaction of transesterication, resulting in a loss of esters as well as causing more fatty
acids to form soaps [143,144].
2.4.3. Reaction temperature
Temperature clearly inuences the reaction and yield of the biodiesel product. A higher reaction temperature can decrease the viscosities of oils and result in an increased reaction rate, and a
shortened reaction time. However, Leung and Guo [28] and Eevera
et al. [143] found that when the reaction temperature increases beyond the optimal level, the yield of the biodiesel product decreases
because a higher reaction temperature accelerates the saponication reaction of triglycerides. The reaction temperature must be
less than the boiling point of alcohol in order to ensure that the
alcohol will not leak out through vaporization. Depending on the
oil used, the optimal temperature ranges from 50 C to 60 C
[28,61,141].
2.4.4. Catalyst concentration
Catalyst concentration can affect the yield of the biodiesel product. As mentioned before, the most commonly used catalyst for the
reaction is sodium hydroxide. However, Freedman et al. [141]
found that sodium methoxide was more effective than sodium
hydroxide because upon mixing sodium hydroxide with methanol
a small amount of water will be produced, which will affect the
product yield because of the hydrolysis reaction [145]. This is the
reason why the catalyst should be added into the methanol rst
and then mixed with the oil. As the catalyst concentration increases the conversion of triglyceride and the yield of biodiesel increase. This is because an insufcient amount of catalysts result in
an incomplete conversion of the triglycerides into the fatty acid esters [28,145]. Usually, the yield reaches an optimal value when the

1091

catalyst (NaOH) concentration reaches 1.5 wt.% and then decreases


a little with a further increase in catalyst concentration. The reduction of the yield of the biodiesel is due to the addition of excessive
alkali catalyst causing more triglycerides to react with the alkali
catalyst and form more soap [28,143].
3. Other processes of biodiesel production
3.1. Biox co-solvent process
The Biox co-solvent process was developed by Boocock et al. in
1996 [146]. In this process, triglycerides are converted to esters
through the selection of inert co-solvents that generate a onephase oil-rich system [86,94,146]. Co-solvent options are available
to overcome slow reaction times caused by the extremely low solubility of the alcohol in the triglyceride phase. Demirbas [94] uses
tetrahydrofuran (THF) as a co-solvent to make the methanol soluble. After the completion of the reaction, the biodieselglycerol
phase separation is clean and both the excess alcohol and the tetrahydrofuran co-solvent can be recovered in a single step [94].
However, because of the possible hazard and toxicity of the co-solvents, they must be completely removed from the glycerol phase
as well as the biodiesel phase and the nal products should be
water-free [146]. The unique advantage of the Biox co-solvent process is that it uses inert, reclaimable co-solvents in a single-pass
reaction that takes only seconds at ambient temperature and pressure, and no catalyst residues appear in either the biodiesel phase
or the glycerol phase [86]. This process can handle not only grainbased feedstocks but also waste cooking oils and animal fats. Van
Gerpen et al. [62] found, however, that the recovery of excess alcohol is difcult when using this process because the boiling point of
the THF co-solvent is very close to that of methanol.
3.2. Supercritical alcohol process
As is known, when a uid or gas is subjected to temperatures
and pressures in excess of its critical point, a number of unusual
properties are exhibited. Under such conditions, a distinct liquid
and vapor phase no longer exist. Instead, a single uid phase is
formed [86]. Therefore, a process for biodiesel production has been
developed by a non-catalytic supercritical methanol method
[147,148]. Because of the lower value of the dielectric constant of
methanol in the supercritical state, this approach is believed to
be able to solve the problems associated with the two-phase nature of normal methanol/triglyceride mixtures by forming a single
phase, and the reaction is completed in a very short time [149].
Supercritical transesterication is carried out in a high pressure
reactor, with heat supplied from an external heater. Reaction occurs during the heating period. After the reaction is complete,
the gas is vented and the product in the reactor is poured into a
collecting vessel. The remaining contents are removed from the
reactor by washing it with methanol [150,151]. During the whole
process, several variables (i.e. reaction pressure and temperature)
affect the yield of the biodiesel product and the highest yield can
be obtained under the optimal conditions. In contrast to the common alkali-catalyzed method, this process has advantages in terms
of reaction time and purication but requires a high temperature
and pressure, hence requiring a high amount of energy
[94,147,148].
3.3. In situ biodiesel process
The in situ biodiesel production is a novel approach for converting oil to biodiesel which was developed by Harrington and DArcyEvans in 1985 [152]. In this method, to achieve transesterication

1092

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095

of its acyglycerols, the oilseeds are directly treated at ambient temperature and pressure with a methanol solution in which the catalyst has been previously dissolved. That means that the oil in the
oilseeds is not isolated prior to transesterication to fatty acid esters [152154]. To reduce the alcohol requirement for high efciency during in situ transesterication, the oilseeds need to be
dried before the reaction takes place [155]. Milled oilseeds are
mixed with alcohol in which the catalyst had been dissolved and
the mixture is heated under reux for 15 h. Two layers are formed
around the time of the completion of the reaction. The lower layer is
the alcohol phase and can be recovered. The upper layer, including
the crude biodiesel, is washed with water to remove the contaminants until the washing solution is neutral. After the washing step,
the upper layer is dried over anhydrous sodium sulfate, then ltered, and the residual product is biodiesel [156]. Haas and Scott
[155] found that the nal biodiesel product can conform to the
ASTM standard and the conversion of the oilseed is very high (about
98%). Since this method eliminates the need for the isolation of, and
possibly for the rening of, the oilseed lipid, the process could reduce biodiesel production costs, reduce the long size of the production system associated with the pre-extraction, degumming, and
maximize the yield of the biodiesel production. However, this process cannot handle waste cooking oils and animal fats, which can
reduce the cost of feedstock [156,157].

There are four primary factors affecting the yield of biodiesel,


i.e. alcohol quantity, reaction time, reaction temperature, and catalyst concentration. To ensure a complete transesterication reaction, the molar ratio of alcohol to triglycerides should be increased
to 6:1 with the use of an alkali catalyst. For used cooking oils or for
oils with a high percentage of free fatty acids, a higher molar ratio
is needed for acid-catalyzed transesterication. Because the conversion rate of fatty acid esters increases with reaction time but
the yield of the biodiesel product reaches a maximum at an optimal reaction time. Higher reaction temperature can decrease the
viscosity of oils, enhancing the reaction rate. The optimal temperature ranged between 50 C and 60 C, depending on the oil used.
The optimal condition of catalyst concentration is about 1.5 wt.%
for NaOH which is the most commonly used catalyst. With increasing concern over global warming, it is foreseeable that biodiesel
usage would continue to grow at a fast pace. This will trigger the
development of more sophisticated methods of biodiesel production and rening to cope with the increasing market demand.
Acknowledgement
The authors would like to acknowledge the support of the ICEE
and the University Development Fund of the University of Hong
Kong.
References

4. Conclusions
Biodiesel is a clean-burning diesel fuel with a chemical structure of fatty acid alkyl esters. Of the various methods available
for producing biodiesel, the alkali-catalyzed transesterication of
vegetable oils and animal fats is currently the most commonly
adopted method. Transesterication is basically a sequential reaction. However, when the raw materials (oils or fats) contain a high
percentage of free fatty acids or water, the alkali catalyst will react
with the free fatty acids to form soaps and the water can hydrolyze
the triglycerides into diglycerides and form more free fatty acids.
These are undesirable reactions which reduce the yield of the biodiesel product. Therefore, after rening the raw materials, the
acidic feedstocks should be pre-treated to inhibit the saponication reaction. There are three primary approaches for reducing
the amount of free fatty acids: esterication of free fatty acids with
methanol, in the presence of acidic catalysts; using iodine as a catalyst; adding glycerol into the acidic feedstock with a catalyst like
zinc chloride and heated to a high temperature. The rst approach
can eliminate the separation, corrosion, toxicity, and environmental problems, but the reaction rate is slower while the advantage of
the last approach is that no alcohol is needed and the water
(formed from the reaction) can be immediately vaporized and
vented from the mixture. The transesterication reaction requires
an alcohol as a reactant and a catalyst. The most commonly used
alcohol is methanol while sodium hydroxide and potassium
hydroxide are the most commonly used catalysts. During the reaction, glycerol will be produced as a by-product. Because of its
numerous industrial applications, the crude glycerol should be rened with purity higher than 99% to make it usable. For rening
the crude biodiesel produced, the product should rst be neutralized and then put through an alcohol stripper before cleaning by
either one of the following approaches: water washing, membrane
extraction, and dry washing. A very promising method for washing
biodiesel is the hollow ber membrane extraction, which effectively avoids emulsication during the washing step and decreases
the rening loss. As a commercial fuel, the nished biodiesel must
be analyzed using sophisticated analytical equipment to ensure
that it meets international standards even if it has been stored
for a long time.

[1] Xu G, Wu GY. The investigation of blending properties of biodiesel and No. 0


diesel fuel. J Jiangsu Polytech Univ 2003;15:168.
[2] Abreu FR, Lima DG, Ham EH, Wolf C, Suarez PAZ. Utilization of metal
complexes as catalysts in the transesterication of Brazilian vegetable oils
with different alcohols. J Mol Catal A Chem 2004;209:2933.
[3] Azcan N, Danisman A. Microwave assisted transesterication of rapeseed oil.
Fuel 2008;87:17818.
[4] Berchmans HJ, Hirata S. Biodiesel production from crude Jatropha curcas L.
seed oil with a high content of free fatty acids. Bioresour Technol
2008;99:171621.
[5] Bernardo A, Howard-Hildige R, OConnell A, Nichol R, Ryan J, Rice B, et al.
Camelina oil as a fuel for diesel transport engines. Ind Crops Prod
2003;17:1917.
[6] Byun MW, Kang IJ, Kwon JH, Hayashi Y, Mori T. Physicochemical properties of
soybean oil extracted from [gamma]-irradiated soybeans. Radiat Phys Chem
1995;46:65962.
[7] Chitra P, Venkatachalam P, Sampathrajan A. Optimisation of experimental
conditions for biodiesel production from alkali-catalysed transesterication
of Jatropha curcus oil. Energy Sustain Dev 2005;9:138.
[8] Demirbas A. Biodiesel fuels from vegetable oils via catalytic and non-catalytic
supercritical alcohol transesterications and other methods: a survey. Energy
Convers Manage 2003;44:2093109.
[9] Kaya C, Hamamci C, Baysal A, Akba O, Erdogan S, Saydut A. Methyl ester of
peanut (Arachis hypogea L.) seed oil as a potential feedstock for biodiesel
production. Renew Energy 2009;34:125760.
[10] Issariyakul T, Kulkarni MG, Meher LC, Dalai AK, Bakhshi NN. Biodiesel
production from mixtures of canola oil and used cooking oil. Chem Eng J
2008;140:7785.
[11] Kansedo J, Lee KT, Bhatia S. Cerbera odollam (sea mango) oil as a promising
non-edible feedstock for biodiesel production. Fuel 2009;88:114850.
[12] Kumar Tiwari A, Kumar A, Raheman H. Biodiesel production from jatropha oil
(Jatropha curcas) with high free fatty acids: an optimized process. Biomass
Bioenergy 2007;31:56975.
[13] Rao Y, Xiang B, Zhou X, Wang Z, Xie S, Xu J. Quantitative and qualitative
determination of acid value of peanut oil using near-infrared spectrometry. J
Food Eng 2009;93:24952.
[14] Sahoo PK, Das LM. Process optimization for biodiesel production from
Jatropha, Karanja and Polanga oils. Fuel 2009;88:158894.
[15] San Jose Alonso J, Lopez Sastre JA, Romero-Avila C, Lopez E. A note on the
combustion of blends of diesel and soya, sunower and rapeseed vegetable
oils in a light boiler. Biomass Bioenergy 2008;32:8806.
[16] Santos FFP, Malveira JQ, Cruz MGA, Fernandes FAN. Production of biodiesel by
ultrasound assisted esterication of Oreochromis niloticus oil. Fuel
2009;doi:10.1016/j.fuel.2009.05.030.
[17] Saraf S, Thomas B. Inuence of feedstock and process chemistry on biodiesel
quality. Process Saf Environ Prot 2007;85:3604.
[18] Singh SP, Singh D. Biodiesel production through the use of different sources
and characterization of oils and their esters as the substitute of diesel: a
review. Renewable and Sustainable Energy Reviews 2009;14:20016.
[19] Srivastava A, Prasad R. Triglycerides-based diesel fuels. Renew Sustain Energy
Rev 2000;4:11133.

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095


[20] Winayanuwattikun P, Kaewpiboon C, Piriyakananon K, Tantong S,
Thakernkarnkit W, Chulalaksananukul W, et al. Potential plant oil feedstock
for lipase-catalyzed biodiesel production in Thailand. Biomass Bioenergy
2008;32:127986.
[21] Zubr J. Oil-seed crop: Camelina sativa. Ind Crops Prod 1997;6:1139.
[22] Schinas P, Karavalakis G, Davaris C, Anastopoulos G, Karonis D, Zannikos F,
et al. Pumpkin (Cucurbita pepo L.) seed oil as an alternative feedstock for the
production of biodiesel in Greece. Biomass Bioenergy 2009;33:449.
[23] Goodrum JW, Geller DP, Adams TT. Rheological characterization of animal
fats and their mixtures with #2 fuel oil. Biomass Bioenergy 2003;24:
24956.
[24] Liu KKM, Barrows FT, Hardy RW, Dong FM. Body composition, growth
performance, and product quality of rainbow trout (Oncorhynchus mykiss)
fed diets containing poultry fat, soybean/corn lecithin, or menhaden oil.
Aquaculture 2004;238:30928.
[25] Patil PD, Deng S. Optimization of biodiesel production from edible and nonedible vegetable oils. Fuel 2009;88:13026.
[26] Gui MM, Lee KT, Bhatia S. Feasibility of edible oil vs. non-edible oil vs. waste
edible oil as biodiesel feedstock. Energy 2008;33:164653.
[27] Haas MJ. Improving the economics of biodiesel production through the use of
low value lipids as feedstocks: vegetable oil soapstock. Fuel Process Technol
2005;86:108796.
[28] Leung DYC, Guo Y. Transesterication of neat and used frying oil:
optimization for biodiesel production. Fuel Process Technol 2006;87:88390.
[29] Upham P, Thornley P, Tomei J, Boucher P. Substitutable biodiesel feedstocks
for the UK: a review of sustainability issues with reference to the UK RTFO. J
Cleaner Prod 2009;17:S3745.
[30] Bhatti HN, Hanif MA, Qasim M, Ataur R. Biodiesel production from waste
tallow. Fuel 2008;87:29616.
[31] Frohlich A, Rice B. Evaluation of Camelina sativa oil as a feedstock for
biodiesel production. Ind Crops Prod 2005;21:2531.
[32] Hameed BH, Lai LF, Chin LH. Production of biodiesel from palm oil (Elaeis
guineensis) using heterogeneous catalyst: an optimized process. Fuel Process
Technol 2009;90:60610.
[33] Liu X, He H, Wang Y, Zhu S, Piao X. Transesterication of soybean oil to
biodiesel using CaO as a solid base catalyst. Fuel 2008;87:21621.
[34] Meher LC, Dharmagadda VSS, Naik SN. Optimization of alkali-catalyzed
transesterication of Pongamia pinnata oil for production of biodiesel.
Bioresour Technol 2006;97:13927.
[35] Ramadhas AS, Jayaraj S, Muraleedharan C. Biodiesel production from high FFA
rubber seed oil. Fuel 2005;84:33540.
[36] Rashid U, Anwar F. Production of biodiesel through optimized alkalinecatalyzed transesterication of rapeseed oil. Fuel 2008;87:26573.
[37] Rashid U, Anwar F, Knothe G. Evaluation of biodiesel obtained from
cottonseed oil. Fuel Process Technol 2009;90:115763.
[38] Rashid U, Anwar F, Moser BR, Ashraf S. Production of sunower oil methyl
esters by optimized alkali-catalyzed methanolysis. Biomass Bioenergy
2008;32:12025.
[39] Demirbas A. Fuel properties and calculation of higher heating values of
vegetable oils. Fuel 1998;77:111720.
[40] Mudge SM, Pereira G. Stimulating the biodegradation of crude oil with
biodiesel preliminary results. Spill Sci Technol Bull 1999;5:3535.
[41] Speidel HK, Lightner RL, Ahmed I. Biodegradability of new engineered fuels
compared to conventional petroleum fuels and alternative fuels in current
use. Appl Biochem Biotechnol 2000;8486:87997.
[42] Harrington KJ. Chemical and physical properties of vegetable oil esters and
their effect on diesel fuel performance. Biomass 1986;9:117.
[43] USEPA. A comprehensive analysis of biodiesel impacts on exhaust emissions.
Draft Technical Report. USEPA; 2002.
[44] Knothe G, Sharp CA, Ryan TW. Exhaust emissions of biodiesel, petrodiesel,
neat methyl esters, and alkanes in a new technology engine. Energy Fuels
2006;20:4038.
[45] Lin L, Ying D, Chaitep S, Vittayapadung S. Biodiesel production from crude rice
bran oil and properties as fuel. Appl Energy 2009;86:6818.
[46] Lopez JM, Gomez A, Aparicio F, Javier Sanchez F. Comparison of GHG
emissions from diesel, biodiesel and natural gas refuse trucks of the City of
Madrid. Appl Energy 2009;86:6105.
[47] Zhang Y, Dube MA, McLean DD, Kates M. Biodiesel production from waste
cooking oil: 2. Economic assessment and sensitivity analysis. Bioresour
Technol 2003;90:22940.
[48] Canakci M, Erdil A, Arcaklioglu E. Performance and exhaust emissions of a
biodiesel engine. Appl Energy 2006;83:594605.
[49] Adams C, Peters J, Rand M, Schroer B, Ziemke M. Investigation of soybean oil
as a diesel fuel extender: endurance tests. J Am Oil Chem Soc 1983;60:
15749.
[50] Anon. Filtered used frying fat powers diesel eet. J Am Oil Chem Soc
1982;59:780A1A.
[51] Engler C, Johnson L, Lepori W, Yarbrough C. Effects of processing and chemical
characteristics of plant oils on performance of an indirect-injection diesel
engine. J Am Oil Chem Soc 1983;60:15926.
[52] Peterson C, Auld D, Korus R. Winter rape oil fuel for diesel engines: recovery
and utilization. J Am Oil Chem Soc 1983;60:157987.
[53] Strayer R, Blake J, Craig W. Canola and high erucic rapeseed oil as substitutes
for diesel fuel: preliminary tests. J Am Oil Chem Soc 1983;60:158792.
[54] Schwab AW, Bagby MO, Freedman B. Preparation and properties of diesel
fuels from vegetable oils. Fuel 1987;66:13728.

1093

[55] Chang C, Wan S. Chinas motor fuels from tung oil. Ind Eng Chem
1947;39:15438.
[56] Crossley A, Heyes TD, Hudson B. The effect of heat on pure triglycerides. J Am
Oil Chem Soc 1962;39:914.
[57] Niehaus RA, Goering CE, Savage Jr LD, Sorenson SC. Cracked soybean oil as
fuel for a diesel engine. Trans Am Soc Agric Eng 1986;29:6839.
[58] Pioch D, Lozano P, Rasoanantoandro MC, Graille J, Geneste P, Guida A. Biofuels
from catalytic cracking of tropical vegetable oils. Oleagineux
1993;48:28991.
[59] Sonntag A. Reactions of fats and fatty acids. In: Swern D, editor. Baileys
industrial oil and fat products. New York: John Wiley & Sons; 1979.
[60] Weisz PB, Haag WO, Rodeweld PG. Catalytic production of high-grade fuel
(gasoline) from biomass compounds by shape-delective catalysis. Science
1979;206:578.
[61] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol
1999;70:115.
[62] Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G. Biodiesel
production technology. 1617 Cole Boulevard, Golden, CO: National
Renewable Energy Laboratory; 2004.
[63] Soriano Jr NU, Venditti R, Argyropoulos DS. Biodiesel synthesis via
homogeneous Lewis acid-catalyzed transesterication. Fuel 2009;88:5605.
[64] ISTC. Feasibility report, small scale biodiesel production. In: Center IST, editor.
Waste Management and Research Center; 2006.
[65] Van Gerpen JH, Dvorak B. The effect of phosphorus level on the total glycerol
and reaction yield of biodiesel bioenergy 2002. In: The 10th Biennial
Bioenergy Conference, Boise; 2002.
[66] Liu GR, Huang YQ, Chen GY. Study on the degumming process of acidic oil. In:
Zhuang X, editor. International conference on biomass energy technologies.
Guangzhou, China; 2008.
[67] de Carvalho CC, de Souza MP, da Silva TD, Gonalves LAG, Viotto LA. Soybeans
crude oil miscella degumming utilizing ceramic membranes: transmembrane
pressure and velocity effects. Desalination 2006;200:5435.
[68] de Souza MP, Cunha Petrus JC, Guaraldo Gonalves LA, Viotto LA. Degumming
of corn oil/hexane miscella using a ceramic membrane. J Food Eng
2008;86:55764.
[69] Verleyen T, Sosinska U, Ioannidou S, Verhe R, Dewettinck K, Huyghebaert A,
et al. Inuence of the vegetable oil rening process on free and esteried
sterols. J Am Oil Chem Soc 2002;79:94753.
[70] Pan L, Noli A, Campana A, Barrera M, Tomas M, Anon M. Inuence of the
operating conditions on acid degumming process in sunower seed oil. J Am
Oil Chem Soc 2001;78:5534.
[71] Eisenmenger M, Dunford N. Bioactive components of commercial and
supercritical carbon dioxide processed wheat germ oil. J Am Oil Chem Soc
2008;85:5561.
[72] Chen YC, He SL, Cheng J. Study on extraction and deacidication of oils and
fats of natural plants by supercritical carbon dioxide. J Yangtze Univ (Nat Sci
Ed) 2007;4:457.
[73] Demirbas A, Kara H. New options for conversion of vegetable oils to
alternative fuels. Energy Sources Part A: Recovery, Utilization &
Environmental Effects. Taylor & Francis Ltd; 2006 [p. 61926].
[74] Mittelbach M, Koncar M. Method for the preparation of fatty acid alkyl esters.
United States Patent 5849939. URL:http://www.freepatentsonline.com/
5849939.html.
[75] Turkay S, Civelekoglu H. Deacidication of sulfur olive oil. I. Single-stage
liquidliquid extraction of miscella with ethyl alcohol. J Am Oil Chem Soc
1991;68:836.
[76] Zhang Y, Lu XH, Yu YL, Ji JB. Study on the coupling process of catalytic
esterication and extraction of high acid value waste oil with methanol. In:
Zhuang X, editor. International conference on biomass energy technologies.
Guangzhou, China; 2008.
[77] Pina C, Meirelles A. Deacidication of corn oil by solvent extraction in a
perforated rotating disc column. J Am Oil Chem Soc 2000;77:5539.
[78] Di Serio M, Tesser R, Dimiccoli M, Cammarota F, Nastasi M, Santacesaria E.
Synthesis of biodiesel via homogeneous Lewis acid catalyst. J Mol Catal A:
Chem 2005;239:1115.
[79] Sheng M, Tian DL, Cao GM. Production of biodiesel fuel from wast edible oil.
China Academic J 2008:26.
[80] Jeromin L, Peukert E, Wollmann G. Process for the pre-esterication of free
fatty acids in raw fats and/or oils. Henkel, Kommanditgesellschaft Auf
Aktien;1990. URL:http://www.freepatentsonline.com/EP0192035B1.html.
[81] Simone MR, Michele CR, Marcelli GR, Paulo LSJ, Fernanda MBC, Elizabeth RL.
Transesterication of vegetable oils promoted by poly(styrenedivinylbenzene)
and
poly(divinylbenzene).
Appl
Catal
A:
Gen
2008;349:198203.
[82] Lou W, Zong M, Duan Z. Efcient production of biodiesel from high free fatty
acid-containing waste oils using various carbohydrate-derived solid acid
catalysts. Bioresour Technol 2008;99:87528.
[83] Ozbay N, Oktar N, Tapan NA. Esterication of free fatty acids in waste cooking
oils (WCO): role of ion-exchange resins. Fuel 2008;87:178998.
[84] Shibasaki-Kitakawa N, Honda H, Kuribayashi H, Toda T, Fukumura T,
Yonemoto T. Biodiesel production using anionic ion-exchange resin as
heterogeneous catalyst. Bioresour Technol 2007;98:41621.
[85] Li C, Liu YH, Luo AX, Ruan RS, Liu CM. New two-step method of producing
biodiesel from waster cooking oil. Cereals Oils Process 2008.
[86] Van Gerpen J, Shanks B, Pruszko R, Clements D, Knothe G. Biodiesel
production technology: August 2002January 2004. Other information:

1094

[87]
[88]
[89]

[90]

[91]
[92]

[93]

[94]
[95]
[96]

[97]

[98]
[99]

[100]

[101]

[102]

[103]

[104]

[105]

[106]

[107]

[108]

[109]
[110]

[111]

[112]

[113]

[114]

[115]

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095


PBD: 1 July 2004; related information: work performed by Iowa State
University, Renewable Products Development Laboratory, and USDA/NCAUR,
2004. Size: 110 p.
Canakci M, Van Gerpen J. Biodiesel production via acid catalysis. Trans Am
Soc Agric Eng 1999;42:120310.
Nelson L, Foglia T, Marmer W. Lipase-catalyzed production of biodiesel. J Am
Oil Chem Soc 1996;73:11915.
Shimada Y, Watanabe Y, Samukawa T, Sugihara A, Noda H, Fukuda H.
Conversion of vegetable oil to biodiesel using immobilized Candida antarctica
lipase. J Am Oil Chem Soc 1999;76:78993.
Devanesan MG, Viruthagiri T, Sugumar N. Transesterication of jatropha oil
using immobilized
pseudomonas uorescens. Afr J Biotechnol
2007;6:2497501.
Fukuda H, Kondo A, Noda H. Biodiesel fuel production by transesterication
of oils. J Biosci Bioeng 2001;92:40516.
Kaieda M, Samukawa T, Matsumoto T, Ban K, Kondo A, Shimada Y, et al.
Biodiesel fuel production from plant oil catalyzed by Rhizopus oryzae lipase
in a water-containing system without an organic solvent. J Biosci Bioeng
1999;88:62731.
Matsumoto T, Takahashi S, Kaieda M, et al. Yeast whole-cell biocatalyst
constructed by intracellular overproduction of Rhizopus oryzae lipase is
applicable to biodiesel fuel production. Appl Microbiol Biotechnol
2001;57:51520.
Demirbas A. Current technologies in biodiesel production. In: Biodiesel; 2008.
p. 16173.
Azcan N, Danisman A. Alkali catalyzed transesterication of cottonseed oil by
microwave irradiation. Fuel 2007;86:263944.
Dias JM, Alvim-Ferraz MCM, Almeida MF. Comparison of the performance of
different homogeneous alkali catalysts during transesterication of waste
and virgin oils and evaluation of biodiesel quality. Fuel 2008;87:35728.
Dizge N, Aydiner C, Imer DY, Bayramoglu M, Tanriseven A, Keskinler B.
Biodiesel production from sunower, soybean, and waste cooking oils by
transesterication using lipase immobilized onto a novel microporous
polymer. Bioresour Technol 2009;100:198391.
Dizge N, Keskinler B. Enzymatic production of biodiesel from canola oil using
immobilized lipase. Biomass Bioenergy 2008;32:12748.
Shimada Y, Watanabe Y, Sugihara A, Tominaga Y. Enzymatic alcoholysis for
biodiesel fuel production and application of the reaction to oil processing. J
Mol Catal B: Enzym 2002;17:13342.
Kulkarni MG, Gopinath R, Meher LC, Dalai AK. Solid acid catalyzed biodiesel
production by simultaneous esterication and transesterication. Green
Chem 2006;8:105662.
Qian WW, Yang L, Lu XP. Preparation of biodiesel catalyzed by KF/CaO with
ultrasonic. In: Zhuang X, editor. International conference on biomass energy
technologies. Guangzhou, China; 2008.
da Silva RB, Lima Neto AF, Soares dos Santos LS, de Oliveira Lima JR, Chaves
MH, dos Santos Jr JR, et al. Catalysts of Cu(II) and Co(II) ions adsorbed in
chitosan used in transesterication of soy bean and babassu oils A new
route for biodiesel syntheses. Bioresour Technol 2008;99:67938.
Di Serio M, Cozzolino M, Tesser R, Patrono P, Pinzari F, Bonelli B, et al. Vanadyl
phosphate catalysts in biodiesel production. Appl Catal A: Gen 2007;320:
17.
Kawashima A, Matsubara K, Honda K. Acceleration of catalytic activity of
calcium
oxide
for
biodiesel
production.
Bioresour
Technol
2009;100:696700.
Lukic I, Krstic J, Jovanovic D, Skala D. Alumina/silica supported K2CO3 as a
catalyst for biodiesel synthesis from sunower oil. Bioresour Technol
2009;100:46906.
Noiroj K, Intarapong P, Luengnaruemitchai A, Jai-In S. A comparative study of
KOH/Al2O3 and KOH/NaY catalysts for biodiesel production via
transesterication from palm oil. Renew Energy 2009;34:114550.
Sakai T, Kawashima A, Koshikawa T. Economic assessment of batch biodiesel
production processes using homogeneous and heterogeneous alkali catalysts.
Bioresour Technol 2009;100:326876.
Shu Q, Zhang Q, Xu G, Nawaz Z, Wang D, Wang J. Synthesis of biodiesel from
cottonseed oil and methanol using a carbon-based solid acid catalyst. Fuel
Process Technol 2009;90:10028.
Zabeti M, Wan Daud WMA, Aroua MK. Activity of solid catalysts for biodiesel
production: a review. Fuel Process Technol 2009;90:7707.
Stavarache C, Vinatoru M, Nishimura R, Maeda Y. Fatty acids methyl esters
from vegetable oil by means of ultrasonic energy. Ultrason Sonochem
2005;12:36772.
Thanh LT, Okitsu K, Sadanaga Y, Takenaka N, Maeda Y, Bandow H.
Ultrasound-assisted production of biodiesel fuel from vegetable oils in a
small scale circulation process. Bioresour Technol 2009;101:63945.
Santos FFP, Rodrigues S, Fernandes FAN. Optimization of the production of
biodiesel from soybean oil by ultrasound assisted methanolysis. Fuel Process
Technol 2009;90:3126.
Wen Z, Yu X, Tu S-T, Yan J, Dahlquist E. Intensication of biodiesel synthesis using zigzag micro-channel reactors. Bioresour Technol 2009;100:
305460.
Schumacher J. Small Scale Biodiesel Production: An Overview. Agricultural
Marketing Policy; 2007. URL:http://www.ampc.montana.edu/policypaper/
policy22.pdf.
Wen GD, Xu YP, Ma HJ, Xu ZS, Tian ZJ. Production of hydrogen by aqueousphase reforming of glycerol. Int J Hydrogen Energy 2008;33:665766.

[116] da Silva GP, Mack M, Contiero J. Glycerol: a promising and abundant carbon
source for industrial microbiology. Biotechnol Adv 2009;27:309.
[117] Wang Z, Zhuge J, Fang H, Prior BA. Glycerol production by microbial
fermentation: a review. Biotechnol Adv 2001;19:20123.
[118] Whittington T. Biodiesel production and use by farmers: is it worth
considering? Western Australia: Department of Agriculture and Food;
2006.
[119] Duncan J. Cost of biodiesel production. Energy Efciency Conserv Auth 2003.
[120] NBB. Biodiesel production and quality. Fuel fact sheets. USA: National
Biodiesel Board; 2007.
[121] Gerpen JV. Biodiesel processing and production. Fuel Process Technol
2005;86:1097107.
[122] Cooke BS, Abrams C, Bertram B. Purication of biodiesel with adsorbent
materials 2005.
[123] Gabelman A, Hwang S. Hollow ber membrane contactors. J Membr Sci
1999;159:61106.
[124] He H, Guo X, Zhu S. Comparison of membrane extraction with traditional
extraction methods for biodiesel production. J Am Oil Chem Soc
2006;83:45760.
[125] Predojevic ZJ. The production of biodiesel from waste frying oils: a
comparison of different purication steps. Fuel 2008;87:35228.
[126] Karaosmanoglu F, Cigizoglu KB, Tuter M, Ertekin S. Investigation of the
rening step of biodiesel production. Energy Fuels 1996;10:8905.
[127] Canakci M, Van Gerpen J. Biodiesel production from oils and fats with high
free fatty acids. Trans Am Soc Agric Eng 2001;44:142936.
[128] Berrios M, Skelton RL. Comparison of purication methods for biodiesel.
Chem Eng J 2008;144:45965.
[129] EN. The EN 14214 standard-specications and test methods; 2003.
[130] ASTM. ASTM D6751-biodiesel blend stock specication (B100).
[131] Mittelbach M. Diesel fuel derived from vegetable oils, VI: specications and
quality control of biodiesel. Bioresour Technol 1996;56:711.
[132] Zheng M, Mulenga MC, Reader GT, Wang MP, Ting DSK, Tjong J. Biodiesel
engine performance and emissions in low temperature combustion. Fuel
2008;87:71422.
[133] Ali Y, Hanna M, Cuppett S. Fuel properties of tallow and soybean oil esters. J
Am Oil Chem Soc 1995;72:155764.
[134] Leung DYC, Koo BCP, Guo Y. Degradation of biodiesel under different storage
conditions. Bioresour Technol 2006;97:2506.
[135] Lee I, Johnson LA, Hammond EG. Use of branched-chain esters to reduce the
crystallization temperature of biodiesel. J Am Oil Chem Soc
1995;72:115560.
[136] Bondioli P, Gasparoli A, Lanzani A, Fedeli E, Veronese S, Sala M. Storage
stability of biodiesel. J Am Oil Chem Soc 1995;72:699702.
[137] Mittelbach M, Gangl S. Long storage stability of biodiesel made from rapeseed
and used frying oil. J Am Oil Chem Soc 2001;78:5737.
[138] Freedman B, Buttereld RO, Pryde EH. Transesterication kinetics of soybean
oil. J Am Oil Chem Soc 1986;63:137580.
[139] Sprules FJ, Price D. Production of fatty esters. United States: NOPCO CHEM
CO; 1950. URL: http://www.freepatentsonline.com/2494366.html.
[140] Zhang D. Crystallization characteristics and fuel properties of tallow methyl
esters. Master thesis, Food Science And Technology. USA: University of
NebraskaLincoln; 1994.
[141] Freedman B, Pryde EH, Mounts TL. Variables affecting the yields of fatty
esters from transesteried vegetable oils. J Am Oil Chem Soc 1984;61:
163843.
[142] Alamu OJ, Waheed MA, Jekayinfa SO, Akintola TA. Optimal transesterication
duration for biodiesel production from nigerian palm kernel oil. Agric Eng
Int: CIGR Ejournal 2007; IX.
[143] Eevera T, Rajendran K, Saradha S. Biodiesel production process optimization
and characterization to assess the suitability of the product for varied
environmental conditions. Renew Energy 2009;34:7625.
[144] Ma F, Clements LD, Hanna MA. The effects of catalyst, free fatty acids, and
water on transesterication of beef tallow. Trans Am Soc Agric Eng
1998;41:12614.
[145] Guo Y. Alkaline-catalyzed production of biodiesel fuel from virgin canola oil
and recycled waste oils. PhD dissertation, Department of Mechanical
Engineering, the University of Hong Kong, Hong Kong; 2005. p. 184.
[146] Boocock DGB, Konar SK, Mao V, Sidi H. Fast one-phase oil-rich processes for
the preparation of vegetable oil methyl esters. Biomass Bioenergy 1996;11:
4350.
[147] Demirbas A. Biodiesel from vegetable oils via transesterication in
supercritical methanol. Energy Convers Manage 2002;43:234956.
[148] Saka S, Kusdiana D. Biodiesel fuel from rapeseed oil as prepared in
supercritical methanol. Fuel 2001;80:22531.
[149] Han HW, Cao WL, Zhang JC. Preparation of biodiesel from soybean oil using
supercritical methanol and CO2 as co-solvent. Process Biochem 2005;40:
314851.
[150] Bunyakiat K, Makmee S, Sawangkeaw R, Ngamprasertsith S. Continuous
production of biodiesel via transesterication from vegetable oils in
supercritical methanol. Energy Fuels 2006;20:8127.
[151] Vera CR, DIppolito SA, Pieck CL, Parera JM. Production of biodiesel by a twostep supercritical reaction process with adsorption rening. In: 2nd Mercosur
congress on chemical engineering, 4th Mercosur congress on process systems
engineering; 2005.
[152] Harrington KJ, DArcy-Evans C. Transesterication in situ of sunower seed
oil. Ind Eng Chem Prod Res Dev 1985;24:3148.

D.Y.C. Leung et al. / Applied Energy 87 (2010) 10831095


[153] Haas M, Scott K, Marmer W, Foglia T. In situ alkaline transesterication: an
effective method for the production of fatty acid esters from vegetable oils. J
Am Oil Chem Soc 2004;81:839.
[154] Siler-Marinkovic S, Tomasevic A. Transesterication of sunower oil in situ.
Fuel 1998;77:138991.
[155] Haas M, Scott K. Moisture removal substantially improves the efciency of
in situ biodiesel production from soybeans. J Am Oil Chem Soc 2007;84:
197204.

1095

[156] Qian JF, Wang F, Liu S, Yun Z. In situ alkaline transesterication of cottonseed
oil for production of biodiesel and nontoxic cottonseed meal. Bioresour
Technol 2008;99:900912.
[157] Mondala A, Liang K, Toghiani H, Hernandez R, French T. Biodiesel production
by in situ transesterication of municipal primary and secondary sludges.
Bioresour Technol 2009;100:120310.

Você também pode gostar