You are on page 1of 9

Bioresource Technology 92 (2004) 297305

Integrated biodiesel production: a comparison


of dierent homogeneous catalysts systems
Gemma Vicente, Mercedes Martnez, Jose Aracil

Department of Chemical Engineering, Faculty of Chemistry, Complutense University, 28040 Madrid, Spain
Received 3 September 2002; received in revised form 6 August 2003; accepted 31 August 2003

Abstract
The most common catalysts for biodiesel production are homogeneous basic catalysts. In the present paper, a comparison is
made of dierent basic catalysts (sodium methoxide, potassium methoxide, sodium hydroxide and potassium hydroxide) for
methanolysis of sunower oil. All the reactions were carried out under the same experimental conditions in a batch stirred reactor
and the subsequent separation and purication stages in a decanter. The analytical methods included gas chromatography and the
determination of fat and oil conventional parameters. The biodiesel purity was near 100 wt.% for all catalysts. However, near 100
wt.% biodiesel yields were only obtained with the methoxide catalysts. According to the material balance of the process, yield losses
were due to triglyceride saponication and methyl ester dissolution in glycerol. Obtained biodiesel met the measured specications,
except for the iodine value, according to the German and EU draft standards. Although all the transesterication reactions were
quite rapid and the biodiesel layers achieved nearly 100% methyl ester concentrations, the reactions using sodium hydroxide turned
out the fastest.
2003 Elsevier Ltd. All rights reserved.
Keywords: Biodiesel; Fatty acid methyl esters; Methanolysis; Transesterication; Sunower oil; Basic catalyst

1. Introduction
Biodiesel is dened as fatty acid methyl or ethyl esters
from vegetable oils or animal fats when they are used as
fuel in diesel engines and heating systems. In this context,
biodiesel shows the following general advantages: (1) An
alternative to petroleum-based fuel, which implies lower
dependence on crude oil foreign imports. (2) Renewable
fuel, helping to achieve the EU renewable energy target (12% of total energy output to consist of renewable
energy by 2010) (European Commission, 1997). (3) A
favourable energy balance. (4) A reduction in greenhouse
gas emissions in line with the Kyoto Protocol agreement.
(5) Lower harmful emissions, which is very advantageous
in environmentally sensitive areas such as large cities and
mines. (6) Biodegradable and non-toxic fuel, being benecial for reservoirs, lakes, marine life and other environmentally sensitive places. (7) The use of agricultural
surplus, as agreed in the European Agricultural Policy
regulations, which can also help to improve rural eco*

Corresponding author. Tel./fax: +91-3944167.


E-mail address: jam1@quim.ucm.es (J. Aracil).

0960-8524/$ - see front matter 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2003.08.014

nomies. Taking the above advantages into consideration,


there is a growing interest in expanding the biodiesel
industry. In this context, the research is focussed on
improving biodiesel quality and yield and increasing the
number of raw materials available.
Fatty acid methyl esters are products of the transesterication (also called methanolysis) of vegetable oils
and fats with methanol in the presence of a suitable
catalyst. In addition, the process yields glycerol. The
reaction scheme is shown in Fig. 1. The stoichiometry of
reaction requires 3 mol of methanol and 1 mol of triglyceride to give 3 mol of fatty acid methyl ester and 1
mol of glycerol. This leads to three consecutive reversible reactions where monoglyceride and diglyceride are
intermediate products. After the reaction, the glycerol is
separated by settling or centrifuging and is puried to be
used in its traditional applications (pharmaceutical,
cosmetics and food industries). In addition, the obtained
glycerol can be used in recently developed applications
in the elds of animal feed, carbon feedstock in fermentations, polymers, surfactants, intermediates and
lubricants (Claude, 1999). The methyl ester phase is also
puried before being used as diesel fuel.

298

G. Vicente et al. / Bioresource Technology 92 (2004) 297305


O

C H2

C
O

C
O

CH

C H2

C H2 OH
O

100 Kg

3 C H3OH

3R

C H3 + C H OH
C H2OH

11 Kg

100 Kg

11 Kg

Fig. 1. Overall scheme of the triglyceride transesterication. The given


amounts are consistent with a molar ratio of methanol:vegetable oil
of 3:1.

The transesterication reaction can be catalysed by


both homogeneous and heterogeneous catalysts. In turn,
the homogeneous catalysts include alkalis and acids.
The most commonly used alkali catalysts are sodium
hydroxide, sodium methoxide and potassium hydroxide.
In this sense, numerous references can be found in the
background literature (Wright et al., 1944; Bradshaw
and Meuly, 1944; Feuge and Gros, 1949; Gauglitz and
Lehman, 1963; Mittelbach et al., 1983; Nye et al., 1983;
Freedman et al., 1984; Lago et al., 1985; Schwab et al.,
1987; Peterson et al., 1991; Fr
ohlich and Rice, 1995a,b;
Boocock et al., 1996; Cvengros and Povazanec, 1996;
Bak et al., 1996; Coteron et al., 1997; Noureddini et al.,
1998; Vicente et al., 1998; Vicente, 2001). Sulphuric acid,
hydrochloric acid and sulfonic acid are usually preferred
as acid catalysts. However, these have been studied less
(Conde Cotes and Wenzel, 1974; Nye et al., 1983;
Freedman et al., 1984; Harrington and DArcy-Evans,
ul and T
1985; Ozg
urkay, 1993; Kildiran et al., 1996;
Siler-Marinkovic and Tomasevic, 1998; Canakci and
Van Gerpen, 1999). Finally, the heterogeneous catalysts
include enzymes (Lazar and Konrad, 1985; Mittelbach,
1990; Nelson et al., 1996; Shimada et al., 1999; Uosukainen et al., 1999), titanium-silicates (Bayense, 1994),
alkaline-earth metal compounds (Peterson and Scarrah,
1984; Gryglewicz, 1999), anion exchange resins (Peterson and Scarrah, 1984) and guanadines heterogenized
on organic polymers (Schuchardt et al., 1996).
The basic catalysts are the most common, since the
process is faster and the reaction conditions are moderated (Reid, 1911; Freedman et al., 1984). However, their
utilisation in vegetable oil transesterication produces
soaps by neutralising the free fatty acid in the oil and
triglyceride saponication. Both soap formations are
undesirable side-reactions, because they partially consume the catalyst, decrease the biodiesel yield and
complicate the separation and purication steps.
The free fatty acid neutralisation can be avoided by
using vegetable oil with a low free fatty acid content
(>0.5%) (Wright et al., 1944; Bradshaw and Meuly,
1944; Feuge and Gros, 1949; Freedman et al., 1984).
However, the most protable raw materials (e.g. waste
cooking oils and fats or low-value fats) usually have a

high content of free fatty acid. Conversely, the saponication side-reaction only takes place when the catalyst
is potassium or sodium hydroxide, because they contain
the necessary hydroxide group (OH) for this reaction.
However, the basic methoxides only have the hydroxide
ion as an impurity. In this sense, they do not produce
soap through trygliceride saponication (Fr
ohlich and
Rice, 1995a).
Likewise, the soap formation can also be avoided by
using an acid catalyst. In addition, the acids catalyse the
free fatty acid esterication to produce fatty acid methyl
esters, increasing the biodiesel yield. Nevertheless, the
acid-catalysed transesterication is much slower than
the basic catalysed reaction and also needs more extreme
temperatures and pressure conditions (Freedman et al.,
1984; Schwab et al., 1987).
More recently, there has been an increase in the development of heterogeneous catalysts to produce fatty
acid methyl esters, because their utilisation in the transesterication reaction greatly simplies and economises
the post-treatment of the products (separation and purication). Besides, the use of heterogeneous catalysts does
not produce soaps through free fatty acid neutralisation
and triglyceride saponication. However, the heterogeneous catalysed reaction also requires extreme reaction
conditions, while the methyl ester yield and the reaction
time are still unfavourable compared to the alkali catalysts (Vicente et al., 1998).
In most studies, the triglyceride conversion rate, the
changes in product composition during reaction or some
of the quality parameters of biodiesel are calculated.
However, only Fr
ohlich and Rice (1995a, 2000) has
dealt with other relevant aspects of the process such as
biodiesel phase yield after the post-treatment stage and
the inuence of the side-reactions on the biodiesel yield.
According to Fr
ohlich, there could be two possible
sources of yield loss in triglyceride transesterication,
namely the dissolution of the methyl ester in the glycerol
phase and the saponication of the triglyceride. The
yield and yield losses can be evaluated through a material balance of the process.
In the present paper, a comparison is made of different basic catalysts for transesterication of sunower
oil. The study is focused on all the above-mentioned
aspects: biodiesel purity and yield, the material balance
of the process, biodiesel quality and changes in product
composition over time.

2. Methods
2.1. Materials
To prepare biodiesel by basic catalysed transesterication, rened sunower oil was obtained from Olcesa
(Cuenca, Spain). The free fatty acid content of the oil

G. Vicente et al. / Bioresource Technology 92 (2004) 297305

was determined according to the AOCS ocial method


(1998) # Ca 5a-40 (0.45 mg KOH/g). The saponication
value was calculated according to the AOCS ocial
method (1998) # Cd 3b-76 (192.2 mg KOH/g). Certied
methanol of 99.8% purity was obtained from Aroca
(Madrid, Spain). The catalystssodium hydroxide,
potassium hydroxide, sodium methoxide and potassium
methoxidewere pure grade from Merck (Barcelona,
Spain). The gas chromatography reference standard for
fatty methyl esters was purchased from Supelco (Madrid, Spain) and for monolein, diolein, and triolein from
Sigma (Madrid, Spain).
2.2. Reaction conditions
All the experiments were carried out under the same
reaction conditions so as to compare the results with the
dierent basic catalysts. The reactions were made at 65
C with a 6:1 molar ratio of methanol to oil and 1% of
basic catalyst by weight of vegetable oil. In turn, each
experiment with a dierent catalyst was repeated three
times to evaluate the experimental errors.
2.3. Equipment
Reactions were carried out in a 100 cm3 three-necked
batch reactor, equipped with a reux condenser, a mechanical stirrer and a stopper to remove samples. The
impeller was set at 600 rpm to avoid mass transfer
limitations on the process. This reactor was immersed
in a constant-temperature bath, which was capable of
maintaining the reaction temperature to within 0.1 C
of the intended gure.
2.4. Experimental procedure
Two types of experimental procedures were carried
out depending on the kind of post-reaction analyses. In
the rst one, the reactor was initially charged with the
desired amount of oil (60 g), then placed in the constanttemperature bath with its associated equipment and
heated to a predetermined temperature. The catalyst was
dissolved in the methanol and the resulting solution was
added to the agitated reactor. The reaction was timed as
soon as the catalyst/methanol solution was added to the
reactor and it continued for 4 h. The mixture was
transferred to a separatory funnel, allowing glycerol to
separate by gravity for 3 h. After removing the glycerol
layer, the methanol was recovered by distillation and the
methyl ester was washed with two volumes of water to
remove catalyst, glycerol and methanol residuals. The
methyl ester and glycerol phases were then analysed to
calculate the biodiesel purity and yield, the material
balance of the process and the quality parameters of
biodiesel.

299

In parallel fashion, each reaction was repeated under


the same experimental conditions in order to determine
the change in transesterication product composition
over time. Samples of 0.5 ml were taken at 2 min intervals and then at longer intervals for 4 h. The samples
were quenched immediately in 0.5 ml of water containing the corresponding amount of HCl to stop the reaction and were subsequently centrifuged. The mixture
was extracted with 1 ml of dichloromethane and the
organic phase was separated, dried with sodium sulphate and evaporated (Mittelbach and Trathnigg, 1990).
The organic layers were then analysed for the quantication of fatty acid methyl esters, monoglycerides, diglycerides and triglycerides.
2.5. Analytical methods
The biodiesel purity means the methyl ester concentration (wt.%) in the biodiesel and is calculated by
capillary gas chromatography. This method also allows
for the quantication of the monoglyceride, diglyceride
and triglyceride contents in the biodiesel (Vicente et al.,
1998). The analyses were performed on a gas chromatograph (Hewlett Packard 5890 Series II, Madrid,
Spain) connected to an integrator (Hewlett Packard
3396SA, Madrid, Spain), using a fused silica capillary
column (Hewlett Packard OV-1, Madrid, Spain) and a
ame-ionization detector (FID).
The biodiesel yield (wt.%), relative to the amount of
vegetable oil poured into the reactor, was calculated
from the methyl ester and vegetable oil weights. In addition, to calculate the material balance of the reactions,
the vegetable oil and the methyl ester and glycerol layers
were analysed. The material balance, which refers to the
initial amount of vegetable oil, includes the molar yield
of biodiesel and the molar yield losses due to triglyceride
saponication and methyl ester in the glycerol phase.
Hence, saponication and acid values were determined
for the vegetable oils and the methyl esters, according to
Cd 3b-76 and Ca 5a-40 AOCS ocial methods (1998),
respectively. Furthermore, 10 g of the glycerol phase
was diluted with 30 ml of water and acidied to a pH of
less than 2 with 3 M sulphuric acid. The mixture was
extracted twice with 20 ml of hexane and once with 20
ml of diethyl ether. The solvents were removed in a
rotary evaporator and the residue was dissolved in 50 ml
of ethanol. Half of the solution was used to determine its
acid value and the remaining half was used to calculate
its saponication value (Fr
ohlich et al., 2000).
The biodiesel quality standards included acid and
iodine values, monoglyceride, diglyceride and triglyceride contents and bonded, free and total glycerine contents. The acid and iodine values were calculated
according to AOCS methods (1998) # Ca 5a-40 and #
Cd 1-25, respectively. The quantication of the monoglyceride, diglyceride and triglyceride contents was

300

G. Vicente et al. / Bioresource Technology 92 (2004) 297305

carried out by capillary gas chromatography, using the


previously described method. The total glycerol level,
expressed as bonded glycerol, was calculated from
the glyceride contents. A photometric analysis based
on enzymatic reactions using commercial test kits
(Boehringer Mannheim, Barcelona, Spain) was used for
analysing free glycerol (Bailer and de Hueber, 1994). In
the end, the total glycerine level was determined from
the previously calculated bonded and free glycerol
content.
Finally, to ascertain the change in the reaction
product composition in the biodiesel phase over time,
the extracted organic phases were analysed by capillary
gas chromatography, which allowed for the simultaneous quantication of fatty acid methyl esters, monoglycerides, diglycerides and triglycerides. The analyses
were made by following the method described in the
previous paragraph.
2.6. Statistical analysis
All the experiments were carried out four times in
order to determine the variability of the results and to
assess the experimental errors. Thus, when the two experimental procedures had been carried out for each
catalyst, both of them were repeated three times. In this
way, the arithmetical averages and the standard deviations were calculated for all the results.

3. Results and discussion


3.1. The Biodiesel purity and yield
To evaluate the biodiesel purity, the methyl ester
concentration (wt.%) in the biodiesel phase was calculated. Conversely, to estimate the biodiesel yield after
the reaction and separation stages, the biodiesel weight
yield, relative to the initial amount of vegetable oil, was
worked out. The results for all the experiments and their
repetitions are shown in Table 1. The arithmetical

averages and standard deviations of the results are also


presented in Table 1. The standard deviations were very
low in all the experiments, indicating a low variation
among the repeated experiments.
When the four catalysts were used, methyl ester
concentrations were nearly 100 wt.%. According to these
results, all the transesterication reactions were completed and, therefore, no dierence in biodiesel purity
was found after 3 h of reaction. However, if there are no
side reactions, the biodiesel weight yields, relative to the
initial amount of vegetable oil, should be nearly 100
wt.%. In this sense, the two possible side reactions are
triglyceride saponication or neutralisation of the free
fatty acid in the vegetable oil. Both of them produce
sodium or potassium soaps and, therefore, decrease the
biodiesel yield. In this case, however, the free fatty acid
neutralisation could not be substantial since the acid
index in the sunower oil was only 0.45 mg KOH/g.
Consequently, triglyceride saponication must be the
only possible side reaction. As shown in Table 1, high
biodiesel yields were obtained by using the sodium or
potassium methoxides (99.33 and 98.46 wt.%, respectively), because they only contain the hydroxide group,
necessary for saponication, as a low proportion impurity. However, when sodium or potassium hydroxides
were utilised as catalysts, biodiesel yields decreased to
86.71 and 91.67 wt.%, respectively. This is due to the
presence of the hydroxide group that originated soaps
by triglyceride saponication. Owing to their polarity,
the soaps dissolved into the glycerol phase during the
separation stage after the reaction. In addition, the
dissolved soaps increased the methyl ester solubility in
the glycerol, an additional cause of yield loss.
3.2. The material balance of the process
In order to quantify these yield losses, the material
balance of the process was determined by analysing the
methyl ester and glycerol phases. The material balance
of the process, relative to the vegetable oil molar
amount, included biodiesel molar yields and also the

Table 1
Eect of the catalyst on the biodiesel purity and yield
Catalyst
Sodium hydroxide

Potassium hydroxide

Sodium methoxide

Potassium methoxide

Biodiesel
purity (wt.%)

99.70
99.75
99.72
99.65

99.71 0.04

99.69
99.80
99.80
99.74

99.76 0.05

99.70
99.69
99.72
99.75

99.72 0.03

99.40
99.50
99.65
99.53

99.52 0.10

Biodiesel
yield (wt.%)

86.33
86.67
87.00
86.71

86.71 0.28

91.67
91.67
91.33
92.00

91.67 0.27

99.17
99.33
99.83
99.00

99.33 0.36

98.33
98.50
98.33
98.67

98.46 0.16

Temperature 65 C, methanol:sunower oil molar ratio 6, catalyst 1 wt.%.

G. Vicente et al. / Bioresource Technology 92 (2004) 297305

301

Table 2
Material balance of the process
Catalyst
Sodium hydroxide

Potassium hydroxide

Sodium methoxide

Potassium methoxide

85.08
85.15
85.22
85.32

85.19 0.10

90.54
90.00
89.67
90.10

98.62
98.78
98.99
98.16

98.64 0.35

97.20
97.66
97.49
97.82

97.54 0.27

Triglyceride
saponication (1)
(molar%)

5.42
5.66
5.66
5.87

5.65 0.18

3.33
3.45
3.48
3.60

3.46 0.11

0.01
0.07
0.05
0.01

0.04 0.03

0.19
0.08
0.10
0.18

0.13 0.05

Methyl ester in
glycerol (2)
(molar%)

6.07
6.03
6.09
5.97

6.04 0.05

2.86
3.07
3.09
2.98

3.00 0.11

0.07
0.12
0.13
0.12

0.11 0.03

0.33
0.48
0.45
0.46

0.43 0.07

Total loss (1) + (2)


(molar%)

11.49
11.70
11.75
11.84

11.69 0.15

6.18
6.53
6.57
6.58

6.46 0.19

0.09
0.19
0.18
0.13

0.15 0.05

0.52
0.56
0.55
0.64

0.56 0.05

Loss not accounted


for (molar%)

3.43
3.15
3.03
2.84

3.11 0.25

3.28
3.48
3.76
3.23

3.44 0.24

1.29
1.02
0.83
1.70

1.21 0.38

2.28
1.78
1.96
1.54

1.89 0.31

Biodiesel yield
(molar%)

90.1 0.36

Temperature 65 C, methanol:sunower oil molar ratio 6, catalyst 1 wt.%.

molar yield losses stemming from triglyceride saponication and methyl ester dissolution in the glycerol phase.
The results, which include the repetitions, the arithmetical averages and the standard deviations, are presented in Table 2. All the calculated standard deviations
were very low. As expected, the yield losses in the reactions with sodium or potassium methoxides were
slight, 0.15 and 0.56 molar%, respectively. However,
when sodium or potassium hydroxide was the catalyst,
there were yield losses due to triglyceride saponication
and methyl ester dissolution in the glycerol phase. As a
result of the lower sodium hydroxide molecular weight
in comparison with the corresponding weight of the
potassium hydroxide, more soaps were produced during
the saponication with sodium hydroxide, 5.65 molar%,
than with potassium hydroxide, 3.46 molar%. This, in
turn, involved an increase in the methyl ester proportion
in the glycerol to 6.04 molar%, when sodium hydroxide

was used as the catalyst. In comparison, the methyl ester


dissolution in the glycerol was only 3.00 molar% after
the reaction with potassium hydroxide.
Nonetheless, when the transesterication was carried
out with nearly equivalent amounts of potassium (1.5
wt.%) and sodium hydroxide (1 wt.%), the quantity of
saponied triglyceride was essentially the same: 5.65 and
5.15 molar%, respectively (Table 3). However, the
amount of methyl ester that dissolved in the glycerol
phase was higher after the reaction with sodium hydroxide (6.04 molar%) than after the reaction with potassium hydroxide (2.18 molar%). At the same time,
after the reaction with sodium hydroxide, it was observed that the subsequent separation stage required
more time because of the lower molecular weight of this
catalyst. Accordingly, this eect was responsible for the
higher amount of methyl ester found in the glycerol
phase.

Table 3
Material balance of the process
Catalyst
Catalyst concentration (wt.%)
Biodiesel yield (molar%)
Triglyceride saponication (1) (molar%)
Methyl ester in glycerol (2) (molar%)
Total loss (1) + (2) (molar%)
Loss not accounted for (molar%)
Temperature 65 C, methanol:sunower oil molar ratio 6.

Sodium hydroxide

Potassium hydroxide

1
85.19 0.10
5.65 0.18
6.04 0.05
11.69 0.15
3.11 0.25

1.5
90.6 0.30
5.15 0.21
2.18 0.10
7.33 0.13
2.07 0.15

302

G. Vicente et al. / Bioresource Technology 92 (2004) 297305

Table 4
Quality control of biodiesel
Catalyst
Sodium
hydroxide
Monoglyceride
content (wt.%)
Diglyceride
content (wt.%)
Triglyceride
content (wt.%)
Bonded
glycerol (wt.%)
Free glycerol
(wt.%)
Total glycerol
(wt.%)
Acid value
(mg KOH/g)
Iodine value
(mg I2 /g)

Potassium
hydroxide

Sodium
methoxide

Potassium
methoxide

EU Draft
standard
prEN 14214

German
standard
DIN51606

US standard
NBB/ASTM

0.295 0.042

0.24 0.053

0.285 0.026

0.48 0.103

Max. 0.8

Max. 0.5

n.d.

n.d.

n.d.

n.d.

Max. 0.2

Max. 0.3

n.d.

n.d.

n.d.

n.d.

Max. 0.2

Max. 0.3

0.076 0.011

0.062 0.014

0.074 0.007

0.124 0.027

Max. 0.2

0.0033 0.0006

0.0032 0.0002

0.0040 0.0008

0.0042 0.0004

Max. 0.02

Max. 0.01

6 0.02

0.0793 0.0107

0.0652 0.0135

0.0780 0.0063

0.1280 0.0270

Max. 0.25

Max. 0.2

6 0.24

0.033 0.004

0.0085 0

0.007 0.001

0.0500 0.002

Max. 0.5

Max. 0.5

Max. 120

Max. 115

128.1 0.7

127.7 0.9

128.5 0.7

128.1 0.8

Temperature 65 C, methanol:sunower oil molar ratio 6, catalyst 1 wt.%.


n.d. not detectable, () no specied limit.

3.3. Quality control of biodiesel


Some of the most important quality parameters of
biodiesel (monoglyceride, diglyceride and triglyceride
content, bonded, free and total glycerol levels, acid value
and iodine value) for reactions using each basic catalyst
are shown in Table 4. These parameters were compared
with some of the standards for biodiesel (The European
Union Draft Standard prEN 14214, The German
Standard DIN 51606 and The United States Standard
NBB/ASTM). The table shows the arithmetical averages
of four experiments and the corresponding standard
deviations. The experimental errors were not signicant
since the standard deviations were very low.
In all cases, the contents of individual glycerides
(monoglycerides, diglycerides and triglycerides) were
within the three specications, which implies that the
transesterication reaction was complete. Consequently,
the bonded glycerol also met the specication parameter. Regarding the free glycerol content, the values
measured were lower than their limited parameter and
this indicated that the glycerol residuals were eliminated
during the purication treatment. As the individual
glyceride and free glycerol levels were under the specications, the total glycerol content also met all the
standards. The acid value is a measure of the fatty acid
level in the biodiesel and depends on the free fatty acid
content in the sunower oil and the transesterication
process. The acid values were within specications in all
reactions.
Concerning the iodine value, it is a measure of the
unsaturation level and, therefore, only depends on the
vegetable oil used as the raw material. In this sense,
sunower oil has high levels of unsaturated fatty acids,

as a result of its high proportion of linolenic acid (three


double bonds). Thus, it registers relatively higher iodine
values in comparison with other raw materials (e.g.
rapeseed oil) with a lower content of unsaturated fatty
acids. For this reason, the measured iodine values were
relatively high (128.1, 127.7, 128.5 and 128.1 mg I2 /g)
and independent of the type of catalyst. Conversely, the
iodine value results were compared with the above
mentioned standards. The iodine values were just
slightly above the recent specied limit, 120 mg I2 /g, in
the draft European Union Standards. However, they
were far above the specication limit, 115 mg I2 /g, according to the German standard. Finally, they met the
United States standard since these specications do not
include the iodine value as a quality parameter. The
dierences among the iodine value specications of
biodiesel are due to the dierent raw materials that have
been used to produce biodiesel in the European Union
and the United States. The countries that rst developed
biodiesel in the European Union (e.g. Germany) utilised
rapeseed oil as a raw material and therefore they included a lower specied limit for the iodine value (115
mg I2 /g). The specied iodine value limit in the European Union draft standard was also initially 115 mg I2 /g,
but it has been changed recently to 120 mg I2 /g to include a large range of raw materials. By contrast, soya
oilwith a similar unsaturation level of sunower oil
has been the most popular raw material in United States
and therefore they do not include a limit for the iodine
value in their standard. In fact, there is not a direct relation between the unsaturation level in the biodiesel
and diesel engine malfunctioning. Nevertheless, the level
of unsaturation is directly linked with the oxidation
tendency. Consequently, in order to avoid oxidation,

G. Vicente et al. / Bioresource Technology 92 (2004) 297305

special precautions must be taken during the storage of


biodiesel from sunower oil. However, methyl esters
form sunower oil have optimal low temperature
properties because of their relatively high level of double
bonds.
3.4. Changes in product composition during transesterication
The methyl ester concentrations (wt.%) in the biodiesel layer over time when using sodium hydroxide,
potassium hydroxide, sodium methoxide and potassium
methoxide are shown in Fig. 2. These results represent
the arithmetical average of four experiments, the standard deviations being lower than 0.50 in all cases. As
can be seen, all the transesterication reactions were
quite rapid and they achieved nearly 100 wt.% methyl
ester concentrations in the biodiesel layer. In this sense,
the reactions that used sodium or potassium hydroxide
were faster than the reactions with the sodium or potassium methoxides. Regarding both types, the transesterication with the sodium catalyst was faster than
the reaction with the potassium catalyst. Thus, the near
100% methyl ester concentration was obtained at 30 min
with sodium hydroxide, at 45 min when using sodium
methoxide or potassium hydroxide and after 4 h with
potassium methoxide.
Transesterication is a three-step reversible reaction.
Each of these steps starts with an attack on the carbonyl
carbon atom of the triglyceride, diglyceride or monoglyceride molecule by the methoxide ion (CH3 O ). For
the potassium and sodium hydroxide, dissolving the
catalyst in methanol produces the methoxide anion:
OH CH3 OH ! CH3 O H2 O
Thus, the quantity of the methoxide anion obtained
depends on the hydroxide ion amount and the alkali
dissociation constant. As sodium and potassium hy-

droxides are strong bases, their dissociation constants


are very high. In this sense, the methoxide ion amount
only depends on the catalyst concentration. The sodium
and potassium catalyst concentrations were the same,
but as sodium hydroxide has a lower molecular weight,
the amount of hydroxide ion needed to produce the
methoxide anion was higher. This situation explains the
faster reaction rate of the transesterication using sodium hydroxide.
Regarding the sodium and potassium methoxides,
their dissociation in methanol produces the necessary
methoxide anion to begin the reaction:
CH3 ONa ! CH3 O Na
Taking into consideration all the catalyst molecular
weights, the amount of methoxide ion in the transesterication with sodium methoxide was lower than the
amount of methoxide produced by using sodium hydroxide. However, it was very similar to the quantity
generated when potassium hydroxide was used. Consequently, the reaction using sodium hydroxide was faster
than the reaction with sodium methoxide and the reaction with the latter was, in turn, faster than the reaction
with potassium hydroxide. Finally, the transesterication with potassium methoxide was the slowest, because
the total methoxide anion quantity was the lowest one.
On the other hand, an additional reaction was carried
out using 1.5 wt.% of potassium hydroxide and repeated
three times to determine the experimental errors. The
arithmetical average results are compared with the corresponding ones for the reaction with 1 wt.% of sodium
hydroxide in Fig. 3. The standard deviations of these
results were lower than 0.5, indicating that the experimental errors were insignicant. The reaction rate was
slightly higher when potassium was used. However, after
10 min, the methyl ester concentration obtained was
essentially the same for both catalysts. This is owing to
the fact that 1 wt.% of sodium hydroxide and 1.5 wt.%

100

100

80
Sodium hydroxide
Potassium hydroxide
Sodium methoxide
Potassium methoxide

60

40

80
Methyl Ester
Concentration
(% wt)

Methyl Ester
Concentration
(% wt)

303

Potassium hydroxide 1.5%


Sodium hydroxide 1%

60

40

20
20
0
0

20

40

60

80 100 120 140 160 180 200 220 240


Time (minutes)

Fig. 2. Change in methyl ester concentration during transesterication. Temperature 65 C, methanol:sunower oil molar ratio 6,
catalyst 1 wt.%.

0
0

10 15 20 25 30 35 40 45 50 55 60
Time (minutes)

Fig. 3. Change in methyl ester concentration during transesterication. Temperature 65 C, methanol:sunower oil molar ratio 6.

304

G. Vicente et al. / Bioresource Technology 92 (2004) 297305


10
9
Methyl Ester /10
Monoglyceride
Diglyceride
Triglyceride /10

Product
Concentration
(% wt)

8
7
6
5
4
3
2
1
0
0

20

40

60

80 100 120 140 160 180 200 220 240


Time (minutes)

Fig. 4. Change in product composition during transesterication.


Temperature 65 C, methanol:sunower oil molar ratio 6, catalyst 1 wt.% KOH.

of potassium hydroxide are nearly equimolecular, but


not fully equimolecular amounts. The equivalent
amount of 1 wt.% of sodium hydroxide is 1.4 wt.% of
potassium hydroxide. Consequently, the quantity of the
hydroxide anion and, therefore, the quantity of the
necessary methoxide anion to start the reaction is
slightly higher when 1.5 wt.% of potassium hydroxide
is utilised as the catalyst.
The product concentration distribution during the
transesterication reaction is presented in Fig. 4 in the
case of potassium hydroxide. This gure shows the arithmetical averages of four experiments, where the corresponding standard deviations were lower than 0.5
and therefore the variability among the repeated experiments was insignicant. The concentrations of intermediatesmonoglycerides and diglyceridesdid not
show a signicant change during the reaction. However,
slight increases in both concentrations were observed
during the rst few minutes of the reaction, achieving
the maximum concentrations, followed by a decrease
to nearly zero, a level maintained until the end of the
reaction. These results indicate that the rst step reaction is the slowest and that it is therefore the ratedetermining reaction.

4. Conclusions
The following conclusions can be drawn from this
study. When the four catalysts were used, the methyl
ester concentrations were near 100 wt.%. High biodiesel
purities are important for biodiesel applications since
residual monoglyceride, diglyceride and triglyceride can
cause serious problems in diesel engines. Biodiesel yields
after the separation and purication steps were higher
than 98 wt.% for the methoxide catalysts, because the
yields losses due to triglyceride saponication and
methyl ester dissolution in glycerol were negligible.

However, these catalysts are more expensive and also


more dicult to manipulate since they are very hygroscopic. The biodiesel yields for the sodium and potassium hydroxide were lower, 85.9 and 91.67 wt.%,
respectively, because the yield losses were more substantial. Nonetheless, these yields can be higher through
a modication of the value for the experimental conditionstemperature and catalyst concentration. The
biodiesel obtained with all the catalysts was within the
European Union draft specications and the established
specications in Germany and United States for
monoglyceride, diglyceride and triglyceride content,
bonded, free and total glycerol levels and acid value. In
this sense, high quality biodiesel was obtained with the
four catalysts. The iodine values were above some of the
standards (The European Union Draft Standard prEN
14214 and The German Standard DIN 51606) because
of the raw material utilisedsunower oil. The transesterication reaction using sodium hydroxide was the
fastest, achieving nearly 100 wt.% methyl ester concentration in the biodiesel phase at 30 min. However, the
reaction rates should be the essentially the same, when
equivalent amounts of the catalysts are used.

Acknowledgements
This work has been funded by the Comisi
on Inteministerial de Ciencia y Tecnologa from Spain (Project CICYT QUI96-0907).

References
AOCS, 1998. Method Cd 3b-76: Saponication Value. In: Firestone,
D. (Eds.), Ocial Methods and Recommended Practices of the
American Oil Chemists Society, fth edition, American Oil
Chemists Society, Champaign, Illinois, USA.
AOCS, 1998. Method Ca 5a-40: Free Fatty Acids. In: Firestone, D.
(Eds.), Ocial Methods and Recommended Practices of the
American Oil Chemists Society, fth edition, American Oil
Chemists Society, Champaign, IL, USA.
AOCS, 1998. Method Cd 1-25: Iodine Value of Fats and Oils. Wijs
Method. In: Firestone, D. (Eds.), Ocial Methods and Recommended Practices of the American Oil Chemists Society, fth
edition, American Oil Chemists Society, Champaign, IL, USA.
Bailer, J.Y., de Hueber, K., 1994. Enzymatical determination of free
glycerol and bonded glycerol. In: Handbook of Analytical Methods for Fatty Acid Methyl Esters used as Diesel Fuel. FITCHTE
Institute, Vienna, pp. 1920.
Bak, Y.-C., Choi, J.-H., Kim, S.-B., Kang, D.-W., 1996. Production of
biodiesel fuels by transesterication of rice bran oil. Kor. J. Chem.
Eng. 13 (3), 242245.
Bayense, C.R., 1994. Esterication process, European Patent no. 0 623
581 A2.
Boocock, D.G.B., Konar, S.K., Mao, V., Sidi, H., 1996. Fast onephase oil-rich processes for the preparation of vegetable oil methyl
esters. Biomass Bioenergy 11 (1), 4350.
Bradshaw, G.B., Meuly, W.C., 1944. Preparation of detergents. US
Patent 2 360 844.

G. Vicente et al. / Bioresource Technology 92 (2004) 297305


Canakci, M., Van Gerpen, J., 1999. Biodiesel production via acid
catalysis. Trans. ASAE 42 (5), 12031210.
Claude, S., 1999. Research of new outlets for glycerol-recent developments in France. Fett/Lipid 101 (3), 101104.
Conde Cotes, A., Wenzel, L., 1974. Cinetica de la transestericaci
on
del aceite de higuerilla. Rev. Lat. Am. Ing. Qum. y Qum. Apl. 4,
125141.
Coteron, A., Vicente, G., Martinez, M., Aracil, J., 1997. Biodiesel
production from vegetable oils. inuence of catalysts and operating
conditions. In: Pandalai, S.G. (Eds.), Recent Res. Developments in
Oil Chemistry 1, Transworld Research Network, India, pp. 109
114.
Cvengros, J., Povazanec, F., 1996. Production and treatment of
rapeseed oil methyl esters as alternative fuels for diesel engines.
Bioresour. Technol. 55, 145152.
European Commission, 1997. Energy for the future: renewable sources
of energy. White Paper for a Community Strategy and Action Plan,
COM (1997) 599 nal of 26/11/1997.
Feuge, R.O., Gros, A.T., 1949. Modication of vegetables oils. VII
Alkali catalyzed interestication of peanut oil with ethanol. J. Am.
Oil Chem. Soc. 26 (3), 97102.
Freedman, B., Pryde, E.H., Mounts, T.L., 1984. Variables aecting the
yields of fatty esters from transesteried vegetable oils. J. Am. Oil
Chem. Soc. 61 (10), 16381643.
Fr
ohlich, A., Rice, B., 1995a. The preparation and properties of
biodiesel grade methyl ester from waste cooking oil. In: Minutes of
the activity meeting of the IEA, Vienna, pp. 1118.
Fr
ohlich, A., Rice, B., 1995b. The preparation and properties of
biodiesel grade methyl ester from Camelina sativa. In: H
oldl, P.,
Schindbauer, H. (Eds.), Proceedings: International Conference
on Standarization and Analysis of Biodiesel. FICHTE, Vienna,
pp. 235240.
Fr
ohlich, A., Rice, B., Vicente, G., 2000. The conversion of waste
tallow into biodiesel grade methyl ester. In: 1st World Conference
and Exhibition on Biomass for Energy and Industry. Proceedings
of the Conference held in Seville (Spain), pp. 695697.
Gauglitz, E.J., Lehman, L.W., 1963. The preparation of alkyl esters
from highly unsaturated triglycerides. J. Am. Oil Chem. Soc. 40,
197198.
Gryglewicz, S., 1999. Rapeseed oil methyl esters preparation using
heterogeneous catalysts. Bioresour. Technol. 70 (3), 249253.
Harrington, K.J., DArcy-Evans, C., 1985. A comparison of conventional and in situ methods of transesterication of seed oil from a
series of sunower cultivars. J. Am. Oil Chem. Soc. 62 (6), 1009
1013.
T
Kildiran, G., Y
ucel, S.O.,
urkay, S., 1996. In-situ alcoholysis of
soybean oil. J. Am. Oil Chem. Soc. 73 (2), 225228.
Lago, R.C., Szpiz, R.R., Jablonka, F.H., Pereira, D.A., Hartman, L.,
1985. Extraction and transesterication of vegetable oils with
ethanol. Oleagineux 40 (3), 147151.
Lazar, V.G., Konrad, H.C., 1985. Estersynthesen mit lipasen. FetteSeifen-Anstrichmittel 10 (1), 394400.

305

Mittelbach, M., 1990. Lipase catalyzed alcoholysis of sunower oil.


J. Am. Oil Chem. Soc. 67 (3), 168170.
Mittelbach, M., Trathnigg, B., 1990. Kinetics of alkaline catalyzed
methanolysis of sunower oil. Fat Sci. Technol. 92 (4), 145148.
Mittelbach, M., W
orgetter, M., Pernkopf, J., Junek, H., 1983. Diesel
fuel derived from vegetable oils: preparation and use of rape oil
methyl ester. Energy Agric. 2, 369384.
Nelson, L.A., Foglia, T.A., Marmer, W.N., 1996. Lipase-catalyzed
production of biodiesel. J. Am. Oil Chem. Soc. 73 (8), 11911195.
Noureddini, H., Harkey, D., Medikonduru, V., 1998. A continuous
process for the conversion of vegetable oils into methylesters of
fatty acids. J. Am. Oil Chem. Soc. 75 (12), 17751783.
Nye, M.J., Williamson, T.W., Deshpande, S., Schrader, J.H., Snively,
W.H., Yurkewich, T.P., French, C.L., 1983. Conversion of used
frying oil to diesel fuel by transesterication: preliminary tests.
J. Am. Oil Chem. Soc. 60 (8), 15981601.
ul, S., T
Ozg
urkay, S., 1993. In situ esterication of rice bran oil with
methanol and ethanol. J. Am. Oil Chem. Soc. 70 (2), 145147.
Peterson, G.R., Scarrah, W.P., 1984. Rapeseed oil transesterication
by heterogeneous catalysis. J. Am. Oil Chem. Soc. 61 (10), 1593
1597.
Peterson, C.L., Feldman, M., Korus, R., Auld, D.L., 1991. Batch type
transesterication process for winter rape oil. Appl. Eng. Agric. 7
(6), 711716.
Reid, E.E., 1911. Studies in esterication. IV. The interdependence
of limits as exemplied in the transformation of esters. Am. Chem.
J. 45, 479516.
Schuchardt, U., Vargas, R.M., Gelbard, G., 1996. Transesterication of
soybean oil catalyzed by alkylguanidines heterogenized on dierent
substituted polystyrenes. J. Mol. Catal. A: Chem. 109 (1), 3744.
Schwab, A.W., Bagby, M.O., Freedman, B., 1987. Preparation and
properties of diesel fuels from vegetables oils. Fuel 66 (10), 1372
1378.
Shimada, Y., Watanabe, Y., Samukawa, T., Sugihara, A., Noda, H.,
Fukuda, H., Tominaga, Y., 1999. Conversion of vegetable oil to
biodiesel using immobilized Candida antarctica lipase. J. Am. Oil
Chem. Soc. 76 (7), 789793.
Siler-Marinkovic, S., Tomasevic, A., 1998. Transesterication of
sunower oil in situ. Fuel 77 (12), 13891391.
Uosukainen, E., Lamsa, M., Linko, Y.-Y., Linko, P., Leisola, M.,
1999. Optimization of enzymatic transesterication of rapeseed oil
ester using response surface and principal component methodology. Enzyme Microbial Technol. 25 (35), 236243.
Vicente, G., 2001. Study of the biodiesel production. PhD Thesis,
Faculty of Chemistry. Complutense University of Madrid.
Vicente, G., Coteron, A., Martnez, M., Aracil, J., 1998. Application
of the factorial design of experiments and response surface
methodology to optimize biodiesel production. Ind. Crops Prod.
8, 2935.
Wright, H.J., Segur, J.B., Clark, H.V., Coburn, S.K., Langdon, E.E.,
DuPuis, R.N., 1944. A report on ester interchange. Oil Soap 21,
145148.

You might also like