You are on page 1of 6

Study on membrane reactors for biodiesel production by phase behaviors

of canola oil methanolysis in batch reactors


Li-Hua Cheng
a,b
, Shih-Yang Yen
b
, Li-Sheng Su
b
, Junghui Chen
b,
*
a
Department of Environmental Engineering, Zhejiang University, Hangzhou 310027, PR China
b
R&D Center for Membrane Technology and Department of Chemical Engineering, Chung-Yuan Christian University, Chung-Li 32023, Taiwan, ROC
a r t i c l e i n f o
Article history:
Received 30 December 2009
Received in revised form 10 March 2010
Accepted 20 March 2010
Keywords:
Biodiesel
Canola oil
Kinetics
Ternary phase diagram
Membrane reactor
a b s t r a c t
In comparison with the general stirring batch reactor, the membrane reactor has been reported to have
higher molar ratios of methanol to oil but ultralow catalyst concentration in the biodiesel production. In
this research, the methanolysis of canola oil is conducted in a stirring batch reactor in the presence of
NaOH as a catalyst. Based on the investigation of the effects of operating conditions, including methanol
to oil molar ration, catalyst concentrations and temperatures, the time course of the reaction path for the
reactant composition in the ternary phase diagram of oilFAMEMeOH offers an effective way to under-
stand the operation of membrane reactors in the biodiesel production. The results show that increasing
the residence time of the whole reactant system within the two-phase zone is good for the separation
operation through the membranes.
2010 Elsevier Ltd. All rights reserved.
1. Introduction
Biodiesel has been widely accepted to be an important compo-
nent of the combined strategic approach to decrease our current
dependence on fossil fuels. Biodiesel is the monoalkyl esters of
long-chain fatty acids made from renewable biological sources,
such as vegetable oils or animal fats. It is also known as fatty acid
methyl ester (FAME). With emissions of less carbon monoxide, sul-
fur compounds, particulate matter and unburned hydrocarbons ex-
cept NO
x
, biodiesel is more advantageous as it is environmental
friendly and non-toxic.
Biodiesel is usually produced by transesterication of vegetable
oils, largely consisting of triglycerides, with methanol. To produce
high purity of biodiesel and to simplify product purication proto-
cols, a membrane reactor has been reported to produce FAME in
the presence of NaOH as a catalyst. The membrane reactor is useful
in removing unreacted canola oil from the FAME product and shift-
ing the reaction equilibrium to the product side (Dube et al., 2007).
The transesterication is believed to occur at the surface of the oil
droplets suspended in methanol, and hence a heterogeneous state
is required to be maintained in the membrane reactor (Cao et al.,
2007). The initial methanol to oil molar ratios of 11:1, 16:1, 23:1
and 46:1 had been tested, and it was reported that at least ratio
of 12:1 was required for the collection of permeate. At a methanol
to oil molar ratio of 24:1, the catalyst concentration can be reduced
to 0.05 wt.% NaOH for the steady-state biodiesel production via the
membrane reactor (Tremblay et al., 2008). With a focus on the
inuences of catalyst concentration and residence time, the kinet-
ics of canola oil transesterication was further investigated by Cao
et al. (2009) in the continuous membrane reactor.
In comparison with high molar ratio of methanol to oil but
ultralow amount of catalyst concentration in the membrane reac-
tor for biodiesel production, molar ratio of 6:1 and catalyst concen-
tration high up to 520 folds are often used in kinetics study in the
stirring batch reactor. Freedman et al. (1986) reported that the
kinetics of the forward reaction, which described the base-cata-
lyzed reaction of methanol and soybean at 6:1 molar ratio, con-
sisted of a combination of second-order consecutive and fourth
order shunt reactions. Noureddini and Zhu (1997) investigated
the kinetics of the transesterication of soybean oil with methanol
using NaOH as a catalyst, and studied the effect of mixing intensity
and temperature on the reaction rates for a 6:1 methanol to soy-
bean oil molar ratio and the catalyst concentration of 0.2 wt.%. A
reaction mechanism was proposed, consisting of an initial mass
transfer-controlled region followed by a second-order kinetically
controlled region. Vicente et al. (2005) conducted a comprehensive
study on the kinetics of sunower oil transesterication with a 6:1
methanol/oil molar ratio, and varied catalyst concentrations of 0.5,
1 and 1.5 wt.%.
Based on the past work, the difference in the methanol to oil
molar ratio and the amount of catalyst concentration between
the membrane reactor and the stirring batch reactor are believed
to be correlated closely with the biodiesel production process, i.e.
0960-8524/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2010.03.095
* Corresponding author. Tel.: +886 3 2654107; fax: +886 3 2654199.
E-mail address: jason@wavenet.cycu.edu.tw (J. Chen).
Bioresource Technology 101 (2010) 66636668
Contents lists available at ScienceDirect
Bioresource Technology
j our nal homepage: www. el sevi er . com/ l ocat e/ bi or t ech
transesterication which is characterized in the reversibility and
mass transfer limitation. The reaction system is essentially hetero-
geneous, as the oil and methanol, with lots of different properties
in their polarity, are almost immiscible in the range from 298.15
to 333.15 K (Cerce et al., 2005). To overcome or rather, exploit this
situation, the two-phase is maintained for the membrane reactor
(Cao et al., 2007, 2009; Dube et al., 2007; Tremblay et al., 2008).
However, as the conversion of the oil, the solubility of oil in meth-
anol gradually increases till nally a homogeneous solution is
formed since the product of FAME can act as a cosolvent (Zhou
et al., 2006). Our previous work also showed that only when the
feed bulk composition was controlled within the two-zone phase,
the oil-rich phase was rejected by the membrane and the permeate
was free of TG (Cheng et al., 2009). Even if the phase behavior of
the typical system composed of oilFAMEMeOH is so important
for affecting the operation performance, the investigation of the
path controlling of the transesterication reaction has not been re-
ported to our best knowledge.
In the present work, the methanolysis of canola oil using NaOH
as catalyst is investigated in a stirring batch reactor. The study is
focused on the inuences of methanol to oil molar ratio and cata-
lyst concentration on the product concentration. The path of the
transesterication reaction within the reactor is characterized with
the liquidliquid phase equilibrium for the system of oilFAME
MeOH. Since the phase diagram is inuenced by the temperature,
the paths of the product compositions in the reactor with the dif-
ferent operating temperatures are studied and compared. For fur-
ther comparison purpose, tting the kinetic model parameters
(Freedman et al., 1986; Vicente et al., 2005), such as the rate con-
stants and the corresponding activation energies, is also carried out
in this study.
2. Methods
2.1. Materials
The canola oil and its biodiesel came from Taiwan NJC Corpora-
tion. Glycerine, monolein, diolein, triolein, butanetriol, tricaprin
and methyl oleate were supplied by SigmaAldrich Co. Methanol
and n-heptane were of HPLC grade. The water content of canola
oil, canola ester and glycerol are 0.05%, 0.03% and 0.1%, respec-
tively. The analysis of the supplied canola ester for the measure-
ment of liquidliquid equilibrium showed that the actual methyl
oleate (C18:1) content was 58.89 wt.%, and the overall content of
C (C16:0C18:3) was 98.68% with TG, DG, MG and free fatty acids
of 0.017%, 0.029%, 0.037% and 0.376%, respectively.
2.2. Equipment
Reactions were carried out in a 250 mL three-necked batch
reactor, where the total volume of reactants was 125 mL. The reac-
tor was equipped with a reux condenser, a thermometer, and a
stopper to remove samples. This reactor was immersed in a con-
stant-temperature bath (Vicente et al., 2005).
2.3. Transesterication
Transesterication conditions employed were shown in Table 1.
The reactor was initially charged with the mixture of canola oil and
methanol, placed in the constant-temperature bath with its associ-
ated equipment, and then heated to a predetermined temperature.
NaOH was dissolved in the minimal amount of methanol, and the
solution was added to the agitated reactor. The reaction was timed
as soon as the NaOH/MeOH solution was added to the reactor, and
it continued for 2 h. During the reaction, samples of 0.5 mL were
taken at the following reaction times: 1, 3, 5, 7, 10, 13, 16, 20,
25, 30, 45, 60, 90 and 120 min. The samples were quenched imme-
diately in the corresponding volume of methanol containing the
oxalic acid of 0.12 g/mL (Zhou, 2006) but not the chloric acid
(Vicente et al., 2005) to stop the reaction. The precipitated sodium
monooxalates were removed by centrifugation of 8000g for
15 min (HsiangTai CN-2200, Taiwan). The organic layers were
evaporated at reduced pressure of 0.337 bar at 60 C using a rotav-
apour (Buchi R-210, Switzerland). The methanol content in the
reaction samples was calculated from the weight loss after evapo-
ration and deduction of the methanol that has been added for the
quenching purpose. The remaining mixture was then analyzed by
capillary gas chromatography (Agilent 6890N) with a ame ioniza-
tion detector (FID) employing an Ultra Alloy
+
-5 capillary column
(Frontier Laboratories LTD, Japan) of 15 m length, 0.53 mm ID,
0.25 lm lmthickness according to ASTM-D6584 (2000) for simul-
taneous determination of components including fatty acid methyl
esters (FAME), monoglycerides (MG), diglycerides (DG), triglycer-
ides (TG), and glycerine (GL). Methyl ester of oleic acid, 1-monoo-
lein, 1,3-diolein, triolein, glycerine were used as standard samples
to determine the retention time of FAME, MG, DG, TG and GL,
respectively. The components, 1,2,4-Butanetriol and tricaprin,
were used as internal standard 1 and 2, respectively. n-Heptane
was used as the solvent for the preparation of GC samples.
2.4. Liquidliquid equilibrium of oilFAMEMeOH
During the transesterication, the time course of three main
components including MeOH, FAME and TG, was characterized
with the ternary phase diagram of oilFAMEMeOH. The ternary
phase diagram of oilFAMEMeOH rather than the intersolubilities
between FAMEMeOHGlycerol, OilGlycerolMeOH and oil
GlycerolFAME, has been used for the kinetics investigation at
the initial stage (Gunvachai et al., 2007). The LLE experiments for
this ternary system were carried out using a water bathed glass
vessel with a magnetic stirrer. The experiments were performed
in a glass vessel of 50 mL volume. Measurements of the boundary
of LLE at atmospheric pressure and different temperatures from 20
to 60 C were carried out by turbidimetric analysis using the titra-
tion method under isothermal conditions (Cerce et al., 2005; Zhou
et al., 2006; Cheng et al., 2009).
2.5. Mathematical modeling for canola oil methanolysis
As shown in Table 1, effects of methanol to oil molar ratio, cat-
alyst concentration and temperature on the reaction rates were fo-
cused in this study. The kinetic modeling approach of Vicente et al.
(2005) was used. Six effective reaction rate constants k
0
1
k
0
6
, six
reaction rate constants k
1
k
6
as shown in formula (1), and the
Table 1
Details of the transesterication experiment.
Methanol to
oil molar ratio
Reaction
temperature (C)
Catalyst
concentration (wt.%)
Run 1 6:1 60 0.1
Run 2 12:1 60 0.1
Run 3 24:1 60 0.1
Run 4 24:1 60 0.5
Run 5 24:1 60 0.05
Run 6 24:1 40 0.1
Run 7 24:1 40 0.5
Run 8 24:1 40 0.05
Run 9 24:1 20 0.1
Run 10 24:1 20 0.5
Run 11 24:1 20 0.05
6664 L.-H. Cheng et al. / Bioresource Technology 101 (2010) 66636668
labels were consistent with those used by Vicente et al. (2005).
They were estimated from measurements performed on the com-
positions of the reactor.
TG MeOH )
*
k
0
1
k
1
C
k
0
2
k
2
C
DG FAME
DG MeOH )
*
k
0
3
k
3
C
k
0
4
k
4
C
MG FAME
MG MeOH )
*
k
0
5
k
5
C
k
0
6
k
6
C
GL FAME
1
where k
0
i
k
i
C, i 1; 2; . . . ; 6, indicating that the effective reaction
constants k
0
i
only depends on the catalyst concentration C and the
rate constants of the catalyzed reaction k
i
.
To calculate the effective rate constants, the minimum differ-
ence (e) between all experiments and calculated concentrations
of the components as shown in Eq. (2) can be used as the objective
criterion of the correctness of the model for base-catalyzed
transesterication.
e min
X
n
i1
X
m
t1
C
exp
i;t
C
fit
i;t

2
2
where n is the number of the reactants, C
i,t
is the concentration of
component i at time t and the superscript exp and t denote
the experimental and tting value, respectively.
3. Results and discussion
In order to fully understand the transesterication reaction path
within the reactor, the time course of the component concentra-
tion including TG, DG, MG, FAME, GL and MeOH was recorded,
and then the composition path was characterized with the ternary
phase diagram of oilFAMEMeOH. Eleven reactions are carried
out by varying the methanol to oil ratio (6:1, 12:1 and 24:1), reac-
tion temperature (20, 40 and 60 C), and catalyst concentration
(0.05, 0.1 and 0.5 wt.% of canola oil).
3.1. Time course of product composition
Fig. 1 shows the product composition variation during the cano-
la oil methanolysis using 1150 rpm at 60 C, 0.1 wt.% of the NaOH,
and methanol to oil molar ratio of 24:1. In this case, the slow rate
region at the initial stage of the reaction is observed even though
the stirring rate of 1150 rpm is adopted. Because of the high molar
ratio of 24:1 and low catalyst concentration of 0.1 wt.%, the immis-
cibility of canola oil in methanol is hard to be overcome as com-
pared with the work of Vicente et al. (2005) where 6:1 of molar
ration and 1.5 wt.% was used, and hence the mass transfer might
still govern the kinetics of the reaction. As a result, increased
mechanical agitation of 1150 rpm is necessary to promote a more
rapid reaction though Vicente et al. found that the maximum value
of impeller speed was 600 rpm (Vicente et al., 2005). Upon vigor-
ous mixing, an emulsion is formed and the reaction is then con-
trolled by the solubility of oil in the catalytically active phase
(Gunvachai et al., 2007). The formation of FAME and byproduct
of glycerol, accompanied by the consumption of methanol and
TG, then increase rapidly until their equilibriums are approached
as shown in Fig. 1, respectively. The concentration of intermediate
MG does not show a signicant change during the whole transeste-
rication reaction. By contrast, about 10 wt.% of the intermediate
DG is detected during the rst few minutes of the reaction, fol-
lowed by a decrease to nearly zero, a level maintained until the
end of the reaction (Vicente et al., 2005). The difference in concen-
tration variation of both DG and MG can be attributed to the high-
est activation energy required for the second forward reaction step
from DG to MG as will be shown in Section 3.3. Although Fig. 1
shows only one of the experimental results, similar time courses
of composition for the other ten reactions have been obtained for
other reaction conditions. Fig. 1 shows that oil, FAME and MeOH
are the three major reaction components in the transesterication
process. Therefore, the intersolubilities of those three components
can be used to represent the main phase variation in the process of
transesterication. This method has also been used in the past re-
search (Gunvachai et al., 2007).
3.2. Effect of kinetics parameters
3.2.1. Effect of methanol to oil molar ratio
Fig. 2 shows the effect of methanol to oil molar ratio on the
composition path during the canola oil methanolysis using
0 20 40 60 80 100 120
0
10
20
30
40
50
60
MeOH / oil = 24:1, T = 60
o
C, NaOH = 0.1 wt.%
Reaction time (min)
C
o
n
c
e
n
t
r
a
t
i
o
n

(
w
t
.
%
)
TG
DG
MG
FAME
GL
MeOH
Fig. 1. Composition of reaction mixture during canola oil methanolysis (methanol
to oil molar ratio = 24:1, temperature = 60 C, NaOH concentration in canola
oil = 0.1 wt.%, stirring speed = 1150 rpm).
0
20
40
60
80
0 20 40 60 80
0
20
40
60
80
5 min
wt.% Canola oil
w
t
.
%

C
a
n
o
l
a

e
s
t
e
r
w
t
.
%

M
e
O
H
1 min
3 min
7 min
10 min
20 min
MeOH / oil = 24:1
MeOH / oil = 12:1
MeOH / oil = 6:1
Experimental LLE at 60
o
C
Fig. 2. Effect of methanol to oil molar ratio on the ternary composition paths of
canola oil methanolysis (T = 60 C, 0.1 wt.% NaOH, stirring speed = 1150 rpm,
samples were taken at reaction times of 1, 3, 5, 7, 10, 13, 16, 20, 25, 30, 45, 60,
90 and 120 min).
L.-H. Cheng et al. / Bioresource Technology 101 (2010) 66636668 6665
1150 rpm at 60 C and 0.1 wt.% of the NaOH. As shown in Fig. 2, oil
and methanol are not mutually soluble so that the initial reaction
composition all starts from the bottom line of MeOHCanola oil,
regardless of the initial methanol to oil molar ratio. However, with
the increase of FAME ratio during the transesterication, the meth-
anol to oil ratio shows signicant effect on the time for the reaction
composition entering into the homogenous phase of the ternary
system of oilFAMEMethanol. If the starting mixture at 60 C
has a molar ratio of 6:1 (open triangle), only 3 min are required
for the reaction to be homogenous. With the increase of methanol
to oil, 10 and 20 min are required before homogenous liquid sys-
tems are formed for molar ratio of 12:1 (open square) and 24:1
(open circle), respectively. This difference in the reaction path ex-
plains partly well for the reasons that in the membrane reactor,
molar ratio of 24:1 is usually adopted for the maintaining of
two-phase status because the oil-rich phase can be rejected by
the membrane barrier, whereas the lower amount of methanol
(like 6:1) is good for the homogenous batch transesterication.
Actually, the effect of the molar ratio on the performance of the
membrane reactor for the biodiesel production has already been
reported experimentally (Cao et al., 2007, 2009; Tremblay et al.,
2008), but no clear explanation was given. In our experiments,
the difference in the reaction path is used to explain why the high-
er molar ratio is usually adopted in the membrane reactor. But
24:1 is not always necessary for the membrane reactor where
the heterogeneous status is required.
3.2.2. Effect of temperature
Fig. 3 shows both the effect of temperatures on the ternary
phase diagram of oilFAMEMeOH, and the composition paths
with the different reaction temperatures during the canola oil
methanolysis using 1150 rpm at molar ratio of 24:1 and 0.1 wt.%
of the NaOH. The liquidliquid equilibrium is apparently inu-
enced by the temperature as shown in Fig. 3. The two-phase area
extends to the higher concentration of the methyl esters at the
lower temperature. For example, the two-phase area at 20 C is
much broader than that at 60 C. Nearly 70 wt.% of methyl ester
is needed to achieve a homogeneous phase at 20 C while only
about 50 wt.% is required at 60 C. It indicates that it is preferable
to operate the membrane reactor under the lower temperature in
order to maintain as broad two-phase area as possible (Cheng
et al., 2009).
However, both the reaction rate and reaction conversion re-
duce when the temperature drops from 60 to 20 C. As shown
in Fig. 3, the nal FAME concentration is about 17 wt.% at the
end of the 2 h of transesterication for the reaction temperature
of 20 C (open circle), whereas nearly 50 wt.% of FAME is obtained
for only 20 min of transesterication at temperature of 60 C
(open triangular). Therefore, in the stirring batch reactor, the reac-
tion temperature is required to be as high as possible for reducing
the reaction time only if the methanol is not evaporated, but for
the membrane reactor, the reaction temperature is required to
be as low as possible for control of two-phase and avoiding the
system at the homogenous phase. Cao et al. had pointed out that
the transesterication was believed to occur at the surface of the
oil droplets suspended in methanol, and heterogeneous phase is
necessary for the operation of the membrane reactor (Cao et al.,
2007). With the increase of the temperature, the nal FAME con-
centration increases. However, the purity of FAME is deteriorated
since the solubility of oil and other intermediates in the product
increase with the temperature. Additionally, our experimental
study as shown in Fig. 3 indicates that the reaction path remained
on the same line irrespective of the increase in reaction tempera-
ture for the constant methanol to oil molar ratio of 24:1 adopted
in this case.
3.2.3. Effect of catalyst concentration
Similar with that shown in Fig. 3, all reaction paths are still
found on the same line as shown in Fig. 4, when the experiments
are performed at 0.05, 0.1 and 0.5 wt.% of NaOH at the constant
molar ration of 24:1. However, the time for the reaction to attain
a steady state reduces with the increase of catalyst concentration
whereas the nal equilibriumconcentration of each component re-
mains almost the same. For the catalyst concentrations of 0.5, 0.1
and 0.05 wt.%, it takes 5, 20 and 45 min for the system to be homo-
geneous as shown on the reaction path line in Fig. 4, and it will take
7, 25 and 60 min to achieve almost the same nal equilibrium
FAME concentration, respectively.
0
20
40
60
80
0 20 40 60 80
0
20
40
60
80
20 min
2 hr
wt.% Canola oil
w
t
.
%

C
a
n
o
l
a

e
s
t
e
r
w
t
.
%

M
e
O
H
20
o
C
40
o
C
60
o
C
Experimental LLE at 20
o
C
Experimental LLE at 40
o
C
Experimental LLE at 60
o
C
Fig. 3. Effect of reaction temperature on the ternary composition paths of canola oil
methanolysis (MeOH:oil = 24:1, 0.1 wt.% NaOH, stirring speed = 1150 rpm, samples
were taken at reaction times of 1, 3, 5, 7, 10, 13, 16, 20, 25, 30, 45, 60, 90 and
120 min).
0
20
40
60
80
0 20 40 60 80
0
20
40
60
80
wt.% Canola oil
w
t
.
%

C
a
n
o
l
a
e
s
t
e
r
w
t
.
%
M
e
O
H
20 min
3
5 min
45 min
Experimental LLE at 60
o
C
NaOH = 0.5 wt.%
NaOH = 0.1 wt.%
NaOH = 0.05 wt.%
Fig. 4. Effect of NaOH concentration on the ternary composition paths of canola oil
methanolysis (MeOH:oil = 24:1, reaction temperature = 60 C, samples were taken
at reaction times of 1, 3, 5, 7, 10, 13, 16, 20, 25, 30, 45, 60, 90 and 120 min).
6666 L.-H. Cheng et al. / Bioresource Technology 101 (2010) 66636668
The corresponding TG conversion in terms of 20 min is calcu-
lated to be 61.1%, 87.2% and 100% for the catalyst concentrations
of 0.05, 0.1 and 0.5 wt.%, respectively. The effect of catalyst on
the concentration of reaction mixture is found apparently different
from the work of Cao et al. using a continuous membrane reactor,
where the overall TG conversion for the 0.05 wt.% catalyst is only
32.7% after 20 min (Cao et al., 2009). The lower TG conversion is
partly attributed to two-phase maintained within the membrane
reactor. Although oil molecules aggregate to form droplets dis-
persed in the alcohol as an emulsion, which is conducive for the
oil-rich phase rejection by the membrane (Cao et al., 2007, 2009),
the mass-transfer effects is still apparent in the membrane reactor
with the methanol to oil molar ratio of 24:1 and the catalyst
concentration of 0.05 wt.%. The mass-transfer resulted from the
two-phase maintenance in the membrane reactor also shows the
necessity to make the transesterication and separation carried
out in two different units, i.e. the transesterication in the stirring
batch reactor as homogenously as possible followed by the
separation in the membrane reactor. Additionally, the same TG
conversion is found to be achieved by the ultralow catalyst concen-
tration of 0.05 wt.% and that of 0.5 wt.% as shown in Fig. 4, indicat-
ing the amount of catalyst can be reduced in this reactor without at
the expense of TG conversion.
To sum up, this above study is to explain why maintaining het-
erogeneous reaction is important to the membrane reactor in the
biodiesel production. Although this has been shown in literature
(Cao et al., 2007, 2009; Tremblay et al., 2008), no strong evidence
has been given to explain why the reaction with the heterogeneous
behavior is good. Although the reactor is related to conventional
reaction, the membrane reactor can simultaneously react and re-
move part of feed and product. It is hard to investigate the phase
behavior directly in the membrane reactor. Fortunately, the reac-
tion mechanism and the phase equilibrium relationships of all
those related components are the same irrespective of the reactor
type. In this research, the methanolysis of canola oil in a stirring
batch reactor is carried out. As a result, the phase behavior ob-
tained from the batch reactor is investigated for the inference of
the operation of the membrane reactor.
3.3. Kinetics for canola oil methanolysis
For comparison purpose, the eleven reactions as shown in
Table 1 are used for the calculation of the effective rate constants
by the mathematical procedure (Vicente et al., 2005), as shown
in Table 2. Similar results can be obtained for the values of the
effective rate constant k
0
6
, corresponding the reverse third reaction,
i.e. it is too small and can be negligible. Also k
0
3
is the highest only
in all the reactions for the methanol to oil molar ratio of 24:1 as
shown in Table 2. When the molar ratio is decreased to 12:1 and
6:1, k
0
4
is slightly larger than k
0
3
. However, the absolute values of
k
0
3
and k
0
4
are about one order lower than the work of Vicente
et al. (2005) in which the batch reactor with molar ratio of 6:1 at
65 C is adopted. By contrast, the order of k
0
3
and k
0
4
are comparable
with that of Cao et al. (2009) where the membrane reactor and the
same methanol to oil ratio are adopted. It shows that the molar ra-
tio of methanol to oil, together with the catalyst concentration has
signicant effect on the apparent reactive constants in the transe-
sterication process.
The effect of the reaction temperature on the rate constant k
1
,
corresponding to the rst forward reaction as shown in Fig. 5, is
found similar to the work of Vicente et al. (2005). That is, k
1
in-
creases with the increase of temperature. Nevertheless, k
1
does
not exceed k
5
, corresponding to the third forward reaction. As
shown in Fig. 5, no cross point is formed for the Arrhenius plot of
k
1
and k
5
. Although all calculated apparent constants increase with
the reaction temperature as shown in Fig. 5, the combined effect of
the molar ratio of methanol to oil and the catalyst concentration
has made the Arrhenius plot of reaction rate versus temperature
signicantly different from that of Vicente et al. (2005).
The temperature dependence of the reaction rate is further cal-
culated as shown in Table 3 for the NaOH-catalyzed reaction by
canola oil. The results show that the activation energies are the
largest for the second step reaction, with slightly higher value for
the forward reaction than that of the reverse reaction. Neverthe-
less, the calculated activation energies for the three-step reactions
are larger than those reported by Vicente et al. (2005), but closer to
those in Freedman et al. (1986) and those in Noureddini and Zhu
(1997). The combined effect of molar ratio of methanol to oil and
Table 2
Apparent rate constants.
k
0
1
(mol/
L min)
k
0
2
(mol/
L min)
k
0
3
(mol/
L min)
k
0
4
(mol/
L min)
k
0
5
(mol/
L min)
k
0
6
(mol/
L min)
Run 1 0.0450 0.0260 0.1052 0.2983 0.1473 0.0006
Run 2 0.0121 0.0556 0.0871 0.3436 0.1934 0.0015
Run 3 0.0084 0.0675 0.3137 0.2836 0.1334 0.0007
Run 4 0.0470 0.3937 1.5660 1.2939 0.6072 0.0038
Run 5 0.0048 0.0277 0.1629 0.1509 0.0908 0.0004
Run 6 0.0039 0.0124 0.0066 0.0042 0.0043 0.0003
Run 7 0.0239 0.0679 0.0341 0.0199 0.0207 0.0018
Run 8 0.0020 0.0062 0.0033 0.0027 0.0015 0.0003
Run 9 0.0004 0.0040 0.0016 0.0016 0.0009 0.0000
Run10 0.0020 0.0218 0.0085 0.0076 0.0062 0.0001
Run 11 0.0002 0.0020 0.0008 0.0008 0.0005 0.0000
2.9 3 3.1 3.2 3.3 3.4 3.5
x 10
-3
-6
-5
-4
-3
-2
-1
0
1
2
3
4
1/T (k
-1
)
l
n

k
k
1
k
2
k
3
k
4
k
5
k
6
Fig. 5. Arrhenius plot of reaction rate versus temperature (MeOH:oil = 24:1, stirring
speed = 1150 rpm).
Table 3
Activation energies and pre-exponential factors.
TG ?DG DG ?TG DG ?MG MG ?DG MG ?FAME FAME ?MG
Activation energies (Ea) (J/mol) 65431.2 58403.2 105093.0 102958.6 92540.5 67587.6
Pre-exponential factors (k
0
) 2.0e + 10 0.9e + 10 5.0e + 17 1.7e + 17 2.2e + 15 3.4e + 9
L.-H. Cheng et al. / Bioresource Technology 101 (2010) 66636668 6667
catalyst concentration causes the difference from those had been
previously reported.
4. Conclusions
From the experimental results, increasing the residence time of
the reaction path within the two-phase zone is inferred to be good
for the simultaneous separation and reaction operations of the
membrane reactor. In fact, in the membrane reactor, part of the
components can be simultaneously removed by the membrane
and the continuous removal of products helps shift the reaction
equilibrium. Thus, the reaction paths for batch and membrane
reactors are not the same. However, if the composition of the sys-
tem can be well controlled, the residence time for the reaction path
within the two-phase zone shown in the phase diagram of oil
FAMEMeOH is longer than that in the conventional batch reactor.
Based on the analysis, enhancing operation performance would be
possible by integrating a pre-reactor before a membrane reactor. It
is being undertaken in our lab.
Acknowledgements
The authors wish to express their sincere gratitude to the Cen-
ter-of-Excellence (COE) Program on Membrane Technology from
the Ministry of Education (MOE), ROC, and to the project Toward
Sustainable Green Technology in the Chung Yuan Christian Univer-
sity, Taiwan, under Grant CYCU-98-CR-CE.
References
ASTM-D6584, 2000. Test Method for Determination of Free and Total Glycerine in
B-100 Biodiesel Methyl Esters by Gas Chromatography. ASTM, USA.
Cao, P., Tremblay, A.Y., Dube, M.A., Morse, K., 2007. Effect of membrane pore size on
the performance of a membrane reactor for biodiesel production. Ind. Eng.
Chem. Res. 46, 5258.
Cao, P., Tremblay, A.Y., Dube, M.A., 2009. Kinetics of canola oil transesterication in
a membrane reactor. Ind. Eng. Chem. Res. 48, 25332541.
Cerce, T., Peter, S., Weidner, E., 2005. Biodiesel-transesterication of biological oils
with liquid catalysts: thermodynamic properties of oilmethanolamine
mixtures. Ind. Eng. Chem. Res. 44, 95359541.
Cheng, L.-H., Cheng, Y.-F., Yen, S.-Y., Chen, J., 2009. Ultraltration of triglyceride
from biodiesel using the phase diagram of oilFAMEMeOH. J. Membr. Sci. 330,
156165.
Dube, M.A., Tremblay, A.Y., Liu, J., 2007. Biodiesel production using a membrane
reactor. Bioresour. Technol. 98, 639647.
Freedman, B., Buttereld, R.O., Pryde, E.H., 1986. Transesterication kinetics of
soybean oil. J. Am. Oil Chem. Soc. 63 (10), 13751380.
Gunvachai, K., Hassan, M.G., Shama, G., Hellgardt, K., 2007. A new solubility model
to describe biodiesel formation kinetics. Trans. I Chem. E, Part B, Process Saf.
Environ. Prot. 85 (B5), 383389.
Noureddini, H., Zhu, D., 1997. Kinetics of transesterication of soybean oil. J. Am. Oil
Chem. Soc. 74 (11), 14571463.
Tremblay, A.Y., Cao, P., Dube, M.A., 2008. Biodiesel production using ultralow
catalyst concentrations. Energy Fuels 22, 27482755.
Vicente, G., Martinez, M., Aracil, J., Esteban, A., 2005. Kinetics of sunower oil
methanolysis. Ind. Eng. Chem. Res. 44, 54475454.
Zhou, H., Lu, H.F., Liang, B., 2006. Solubility of multicomponent systems in the
biodiesel production by transersterication of Jatropha curcas L. oil with
methanol. J. Chem. Eng. Data 51, 11301135.
Zhou, W.Y., 2006. Kinetics and Phase Behavior of Transersterication of
Triglycerides. Ph.D. Thesis, University of Toronto, Canada.
6668 L.-H. Cheng et al. / Bioresource Technology 101 (2010) 66636668

You might also like