Advances in Chemical Engi neering and Science , 20 1 1, 1, 45-50
doi:10.4236/aces.2011.12008 Published Online April 2011 (http://www.scirp.org/journal/aces)
Copyright © 2011 SciRes. ACES
Effects of Molar Ratio, Alkali Catalyst Concentration and
Temperature on Transesterification of Jatropha Oil with
Methanol under Ultrasonic Irradiation
Amish P. Vyas, Jaswant L. Verma, Nandula Subrahmanyam
Chemical Engineering Department, Nirma University, Ahmedabad, India
E-mail: Amish71in@yahoo.com
Received December 28, 2010; revised March 10, 2011; accepted March 21, 2011
Abstract
For transesterification of Jatropha oil into biodiesel, ultrasound assisted transesterification seems to be prom-
ising in terms of reduction in process time and stages of operation. Effects of process variables such as the
catalyst loading, the molar ratio of methanol to oil, reaction temperature and the reaction time were investi-
gated on the conversion of Jatropha oil to biodiesel. The conversion was above 93% under the conditions of
50˚C, methanol to oil molar ratio of 9:1, reaction time 30 min and catalyst amount (catalyst/oil) of 1% wt%.
A kinetic study of transesterification of Jatropha oil based on ultrasound assisted synthesis is presented in
this paper. Rate equation obtained is also presented.
Keywords: Transesterification; Ultrasound; Triglyceride; Jatropha Oil, Alkali Catalyst
1. Introduction
Vehicular pollutant emissions caused by the combustion
of fossil fuels and crude oil price fluctuations brought
into focus the need for developing alternate fuels which
could create less pollution, produced from renewable
feedstocks and operate without much modification in the
existing design of the engine. Biodiesel (fatty acid alkyl
ester) derive d f rom transesteri fication of vegetable oils or
an animal fat with methanol (Figure 1) is a potential
substitute for petroleum based diesel fuels. Even 5% re-
placement of petroleum fuel by biofuel can save a coun-
try like India Rs. 4000 crores per year in foreign ex-
change [1].
Government of India has already given due impor-
tance to biofuel and announced a National Biofuel policy
in year 2006. The focus is on collection and distribution
of renewable feedstocks for biofuel products and R & D
at pilot plant scale and later scaling upto commercial lev-
Figure 1. Transesterification of triglycerides with alcohol.
el technologies for production of biodiesel using Jatropha
oil and Karanja oil. At present, the biodiesel is usually
produced by reacting methanol and a vegetable oil in a
batch stirred tank reactor using a liquid alkaline catalyst.
Ultrasound assisted transesterification process offers a
number of advantages over current technology, namely
the simplification of the process and downstream separa-
tion. The present study involved transesterification of
Jatropha oil with methanol catalyzed by alkali catalyst.
Effects of various parameters were studied.
2. Ultrasound Technology
Influence of ultrasound on transesterification reaction is
of purely physical nature. Formation of fine emulsion
between oil and al cohol due t o microt urbulence ge nerated
by cavitation bubbl es generates enormous interfacial area,
which accelerates the reaction [2]. Ultrasound is the
process of propagation of the compression waves with
frequencies above the range of human hearing. Ultra-
sound frequency ranges from 20 kHz to l0 MHz, with
associated acoustic wavelengths in liquids of about 100-
0.15 mm. These wavelengths are not on the scale of mo-
lecular dimensions. Instead, the chemical effects of ul-
trasound derive from several nonlinear acoustic phe-
nomena, of which cavitation is the most important.
Acoustic cavitation is the formation, growth, and implo-
A. P. VYAS ET AL.
Copyright © 2011 SciRes. ACES
46
sive collapse of bubbles in a liquid irradiated with sound
or ultrasound. When sound passes through a liquid, it
consists of expansion (negative pressure) waves and
compression (positive pressure) waves. These cause
bubbles (which are filled with both solvent and solute
vapour and with previously dissolv ed gases) to grow and
recompress. Under proper conditions, acoustic cavitation
can lead to implosive co mpression in such cavities. Such
implosive bubble collapse produ ces inten s e local h eating,
high pressures, and very short life-times. Cavitation is an
extraordinary method of concentrating the diffused en-
ergy of sound into a chemically useab le form. Ultrasoni-
cation provides the mechanical energy for mixing and the
required activation energy for initiating the transesterifi-
cation reaction. Low-frequency ultrasonic irradiation is
useful tool for emulsification of immiscible liquids. The
collapse of the cavitation bubbles disrupts the phase
boundary and causes emulsification, by ultrasonic jets
that impinge one liquid on another [3,4].
Effect of low -frequency ultrasound was studied on the
production of biodiesel v ia transesterification of Jatropha
oil with methanol using sodium hydroxide as homoge-
neous catalyst.
3. Experimental Work
3.1. Reagents and Materials
Methanol (99.5%) and sodium hydroxide (98%) pur-
chased from Yash Enterprise, Ahmedabad, Gujarat, India.
Jatropha oil was purchased from Nidhita marketing,
Ahmedabad, Gujarat, India. Properties of purchased Jat-
ropha oil as carried out at Nirma University are shown in
Table 1.
3.2. Transesterification Procedure
Methanol and Jatropha oil were used as a raw material to
study the eff ect of low fr equency ultr asound on biodiesel
production at 303 K, 313 K, 323 K using molar ratio of
oil to methanol ranging from 1:3, 1:6, 1:9 and 1:12, and
the quantity of alkali catalyst from 0.5%, 1.0%, 1.5%,
and 2% (wt/wt) of the weight of Jatropha oil. The reac-
tion mixture consists of Jatropha oil, methanol, and so-
dium hydroxide. Sodium hydroxide was dissolved into
methanol followed by addition of Jatropha oil to the so-
lution. Since, the Jatropha oil and methanol were not
Table 1. Properties of Jatropha oil.
Property Value
Acid value (mg KOH/gm) 6.171
Saponification value (mg/ g) 198.5
Iodine value (mg iodine/g) 227
Free fatty acid 3.0855
Viscosity (mm2/s) 47
completely miscible, two layers were observed: the up-
per layer was of methanol and the lower layer was of oil.
Reaction mixture was taken in conical flask and it is
placed into water bath of Toshniwal Ultrasonic Cleaner
(30 kHz frequency), which was used for ultrasonication,
after achieving desired reaction temperature of water in
ultrasonic cleaner. During reaction under ultrasound ir-
radiation, no external stirring is provided. The mixing
took place due to the effect of ultrasound wave mecha-
nism in water-bath. The reaction mixture was kept for
the predecided reaction time under ultrasonic effect and
then taken for centrifugation at 1800 to 2000 rpm using
laboratory centrifuge. Later the mixture was allowed to
settle in separating funnel to get two separate layers of
biodiesel and glycerine. Separation of biodiesel and
glycerine was done using a separating funnel. Water
wash was given to the separated biodiesel to remove wa-
ter soluble impurities and then dried to remove moisture
by putting under calcination.
4. Results and Discussion
4.1. Properties of Biodiesel
Biodiesel produced was analyzed and compared with the
fuel properties of petro-diesel as per ASTM Standards
(Table 2). The Jatr opha oil, however, was found to have
much higher values of fuel properties, especially kine-
matic viscosity (Table 1), restricting the direct use as a
fuel in diesel engine. After transesterification, the kine-
matic viscosity value reduced to permissible limit. The
flash points were in a limit of safe storage and handling
conditions. The pour point and cloud points were little
higher than petro-diesel. This might be due to presence
of wax, which begins to crystallize with the decrease in
temperature. This problem could be solved by using
blend of biodiesel with petro-diesel.
4.2. Influence of Reaction Parameters
Experiments were carried out by changing different
process parameters.
Table 2. Comparison fuel properties of petro-diesel, bio-
diesel and ASTM D6751.
Fuel properties Petro-diesel Biodiesel
(B100) ASTM D6751
Specific gravity at
15°C 0.846 0.875 0.86 - 0.90
Kinematic viscosity
(mm2/s) at 40 °C 1.9 - 6.0 4.756 4 - 6
Flash Point (°C) Minimum 1 30 165 100 - 170
Cloud Point (°C) –5 –3 –3 to 12
Pour Point (°C) –10 –8 –15 to 10
Cetane number Minimum 47 54.25 48 - 65
A. P. VYAS ET AL.
Copyright © 2011 SciRes. ACES
47
4.2.1. Methanol to Oil Molar Ratio
Stoichiometrically, the methanolysis of Jatropha oil re-
quires three moles of methanol for each mole of oil.
Since, transesterification of triglycerides is reversible
reaction; excess methanol is requ ired to shift the equilib-
rium towards the direction of ester formation. As can be
seen from Figure 2 (at 303 K) and Figure 3 (at 323 K),
the maximum conversion was achieved at methanol to
oil molar ratio 9:1. It is comparable to the work carried
out by H. D. Hanh et al. [5] obtained 90% conversion
using methanol as an alcohol with triolein oil to alcohol
molar ratio of 1:6 and KOH as a catalyst. D. Kumar et. al.
[6] have obtained above 98% yield using 1:9 Jatropha oil
to methanol molar ratio and heterogeneous solid catalyst
used was Na/SiO2. Present study shows that with molar
ratio of oil to methanol of 1:12, maximum conversion
was achieved in 30 minutes only and after that it almost a
constant over an extended reaction time. Molar ration of
1:3 and 1:6 are not showing good results. One of the
reasons for the same may be the predominance of esteri-
fication reaction at the initial phase, to transesterify the
FFA present in the Jatropha oil, of transesterification
which can consume methanol present in the reaction
mixture and hence, the amount of methanol available for
transesterification may not be sufficient to drive the reac-
tion forward for longer time.
4.2.2. Amount of Catalyst
Effect of variation of amount of catalyst on conversion
was also studied. Catalyst amount was varied in the
range of 0.5% to 2.5% (wt/wt of the oil taken). As shown
in Figure 4, the conversion increased firstly with the
increase of catalyst amount from 0.5% to 1.5%. But, with
further increase in the catalyst amount from 1.5% to
2.5%, the convers ion decreased due to soap for mation. D.
Kumar et al. [6] obtained their best result at 3% wt%
catalyst amount which is higher than the present study.
Separation of heterogeneous catalyst is adding one more
stage in the process presented by D. Kumar. H. D. Hanh
et al. [5] obtained about 90% conversion with 1% wt%
of NaOH catalyst. The conversion obtained by them is
less than what obtained in present study using same
amount of catalyst.
4.2.3. Reaction Time and Temperature
Figure 5 shows the conversion versus reaction time at
different temperatures. It could be seen from the plot that
the conversion increased in the reaction time range of 10
to 45 minutes with the increase in tempera ture, and th ere
after remained nearly constant as a representative of a
nearby equilibrium conversion. The nearly equilibrium
conversion was found to be about 93.5% at 45 minutes of
reaction time. Effect of reaction temperature is not stud-
ied by D. Kumar et al. [6] using heterogeneous solid
Figure 2. Effect of molar ratio on conversion of triglyceride
at 303 K. Reaction conditions: Jatropha oil 50 g, catalyst
amount 1%.
Figure 3. Effect of molar ratio on conversion of triglyceride
at 323 K. Reaction conditions: Jatropha oil 50 g, catalyst
amount 1%.
Figure 4. Effect of amount of catalyst on triglyceride
conversion. Reaction conditions: Jatropha oil 50 g, Oil to
Methanol Molar Ratio 1:9, Reaction time 30 min for 323 K
and 45 min for 303 K.
catalyst but results published (Table 5) for feedstocks
other than Jatropha oil shows that reaction time obtained
in present study is at par w ith other published results.
5. Reaction Kinetics
Experiments were carried out at three different tempera-
A. P. VYAS ET AL.
Copyright © 2011 SciRes. ACES
48
Figure 5. Effect of reaction temperature on triglyceride
conversion. Reaction conditions: Jatropha oil 50 g, catalyst
amount 1%, Oil to Methanol Molar Ratio 1:9.
tures 303 K, 313 K and 323 K. It could be seen from
Figure 5 that with increase in reaction temperature con-
version also increased. The overall rate equation (k) can
be found using (1) for first order reaction and (2) for
second order reaction.

ln 1t
Α
kX  (1)


ln 1t
Ao ΑΑ
kCX X (2)
where XA is the conversion of triglyceride.
The experimental data collected at 303 K, 313 K and
323 K were tested for 1st order kinetics (Figures 6 to 8
and Ta ble 3) and 2nd order kinetics (Figures 9 to 11, and
Table 3). The experimental data fitted well for 1st order
kinetics. The rate data collected were subjected to Ar-
henius equation as follows:

0exp kk ΕRΤ (3)

0
ln ln kΕRΤk (4)
Figure 6. Plot of –ln(1 – XA) vs time, t (minutes) at 323 K
assuming first order reaction. Rate constant k = 0.091 min-1.
Oil: Methanol = 1:9, NaOH = 1% wt%.
Table 3. Reaction rate constant k (min–1) at different tem-
peratures for first order and second order reactions as-
sumption, respectively.
Order of
Reaction
(Assumed)
Reaction
Temperature,
K Rate Constant, k R
2
First 303 0.06 min–1
0.944
313 0.075 min–1
0.967
323 0.091 min–1
0.974
Second 303 353.10 ml/mol*min–1 0.807
313 447.96 ml/mol*min–1 0.807
323 534.92 ml/mol*min–1 0.807
Figure 7. Plot of –ln(1 – XA) vs time, t (minutes) at 313 K
assuming first order reaction. Rate constant k = 0.075 min–1.
Oil: Methanol = 1:9, NaOH = 1% wt%.
Figure 8. Plot of –ln(1 – XA) vs time, t (minutes) at 303 K
assuming first order reaction. Rate constant k = 0.06 min–1.
Oil: Methanol = 1:9, NaOH = 1% wt%.
where, E is activation energy, R is the gas constant
(J/mol-K), T is an absolute temperature, and k0 is a fre-
quency facto r .
From the plot of lnk versus 1/T (Figure 12, Table 4),
–E/R = Slope of the graph = –2039. Therefore, Activa-
tion energy E = 2039 × 1.987 = 4051.49 cal/mol, and
lnk0 = 3.918
k0 = 50.30
A. P. VYAS ET AL.
Copyright © 2011 SciRes. ACES
49
Table 4. Database for determination of activation energy.
Reaction
Temperature, T
(K)
1/T Rate Constant, k
(min–1)
lnk
303 0.0033 0.06 min–1
–2.81341
313 0.0031950.075 –2.59027
323 0.0030960.091 –2.3969
Rate equation derived from the above data of activa-
tion energy for the ultrasound assisted transesterification
reaction is:

dd 1
AAAo Α
rXtkC X (5)
where,

0expkk ΕRΤ 
Substituting the values of k0, E and R in (3) and also
substituting (3) in (5) the rate equation for the reaction is:


50.30exp4051.49 1.9871
AAo Α
rC TX 
(6)
6. Conclusions
Ultrasound assisted transesterification reaction for the
production of biodiesel was found to be very promising
from the results obtained. The optimum conditions for the
production of biodiesel from Jatropha oil under the full
ultrasound condition were molar ratio of oil to methanol
of 1:9 with NaOH concentration of 1 wt% and reaction
time of 30 min. Increasing reaction time and temperature
as well as the molar ratio contributes to high conversion
of triglyceride. The transesterification of Jatropha oil un-
der ultrasound condition provides a possibility for pro-
ducing cheap alternative fuels, which will reduce pollu-
tion and protect the environment. The conversion is com-
parable to the results reported by C. Stavarache et al. [7]
Figure 9. Plot of –ln[XA/(1 – XA)] vs time, t (minutes) at 323
K assuming second order reaction. CAo = 0.00075899 mol/ml.
Rate constant k = 534.92 ml/mol*min–1. Oil: Methanol = 1:9,
NaOH = 1% wt%.
Figure 10. Plot of –ln[XA/(1 – XA)] vs time, t (minutes) at 313
K assuming second order reaction. CAo = 0.00075899 mol/ml.
Rate constant k = 447.96 ml/mol*min–1. Oil: Methanol = 1:9,
NaOH = 1% wt%.
Figure 11. Plot of –ln[XA/(1 – XA)] vs time, t (minutes) at 303
K assuming second order reaction. CAo = 0.00075899 mol/ml.
Rate constant k = 353.10 ml/mol*min–1. Oil: Methanol = 1:9,
NaOH = 1% wt%.
Figure 12. Plot of lnk vs 1/T for the transesterification reac-
tion under ultrasound conditions.
where they used n-propanol as a solvent which is higher
molecular weight alcohol compared to methanol used in
this study. Other reported result was for the feedstock
A. P. VYAS ET AL.
Copyright © 2011 SciRes. ACES
50
Table 5. Comparative study of ultrasound assiste d tr anse sterification [8].
Oil/Tri-
olein Catalyst
Catalyst
wt% of
oil
Alcohol
Oil to
Alcohol
Molar
Ratio
Ultrasonic
Frequency
Source of
Ultrasound
Reaction
Conditions
Ester
Yield, %
Ester
Conver-
sion, %
Ref.
Triolein KOH 1 Methanol 1:6 40 kHz Ultrasonic Cleaner
(1200 W)
25,
10 min - > 90 [5]
NA NaOH 0.5 n-Propanol 1:6 28 kHz
40 kHz
Ultrasonic Cleaner
(1200 W)
25,
20 min
25,
20 min
92
88
-
- [7]
Triolein NaOH 1 Ethanol 1:6 40 kHz
Ultrasonic Cleaner
(1200 W)
25,
< 20 min - 98 [9]
Soy
b
ean
frying
Oil
NaOH 1.5 Methanol NA 24 kHz Ultrasonicator
(200 W)
60,
20 min 97 - [10]
Fish Oil C2H5ONa 0.8 Ethanol 1:6 20 kHz Ultrasonic Probe
60,
60 min 98.2 - [8]
of soybean frying oil [10]. The present study compares
favourably with the results reported in literature. Other
comparable published results of different researchers are
presented in Table 5. Since, the process seems economic;
the economic st udy i s t o be carri ed o ut .
7. Acknowledgements
The authors thank the authorities of the Nirma Univesity,
Ahmedabad for supporting the investigation.
8. References
[1] A. P. Vyas, N. Subrahmanyam and P. A. Patel, “Produc-
tion of Biodiesel through Transesterification of Jatropha
Oil Using KNO3/Al2O3 Solid Cataly st,” Fuel, Vol. 88, No.
4, 2009, pp. 625-628. doi:10.1016/j.fuel.2008.10.033
[2] A. Kalva, T. Sivasankar and V. S. Moholkar, “Physical
Mechanism of Ultrasound-Assisted Synthesis of Bio-
diesel,” Industrial & Engineering Chemistry Research,
Vol. 48, No. 1, 2009, pp. 534-544.
doi:10.1021/ie800269g
[3] A. K. Singh, S. D. Fernando and R. Hernandez, “Base-
Catalyzed Fast Transesterification of Soybean Oil Using
Ultrasonication,” Energy & Fuels, Vol. 21, No. 2, 2007,
pp. 1161-1164. doi:10.1021/ef060507g
[4] A. P. Vyas, J. L. Verma and N. Subrahmanyam, “A Re-
view on FAME Production Processes,” Fuel, Vol. 89, No.
1, 2010, pp. 1-9. doi:10.1016/j.fuel.2009.08.014
[5] H. D. Hanh, N. T. Dong, K. Okitsu, R. Nishimura and M.
Yasuaki, “Effects of Molar Ratio, Catalyst Concentration
and Temperature on Transesterification of Triolein with
Ethanol under Ultrasonic Irradiation,” Journal of the Ja-
pan Petroleum Institute, Vol. 50, No. 4, 2007, pp. 195-
199. doi:10.1627/jpi.50.195
[6] D. Kumar, G. Kumar and C. P. P. Singh, “Ultrasonic-
Assisted Transesterification of Jatropha curcus Oil Using
Solid Catalyst, Na/SiO2,” Ultrasonics Sonochemistry, Vol.
17, No. 5, June 2010, pp. 839-844.
doi:10.1016/j.ultsonch.2010.03.001
[7] C. Stavarache, M. Vinatoru, R. Nishimura and Y. Maeda,
“Fatty Acids Methyl Esters from Vegetable Oil by Means
of Ultrasonic Energy,” Ultrasonics Sonochemistry, Vol.
12, No. 5, 2005, pp. 367-372.
doi:10.1016/j.ultsonch.2004.04.001
[8] R. E. Armenta, M. Vinatoru, A. M. Burja, J. A. Kralovec
and C. J. Barrow, “Transesterification of Fish Oil to Pro-
duce Fatty Acid Ethyl Esters Using Ultrasonic Energy,”
Journal of the American Oil Chemists Society, Vol. 84,
No. 11, 2007, pp. 1045-1052.
[9] H. D. Hanh, N. T. Dong, K. Okitsu, R. Nishimura and M.
Yasuaki, “Biodiesel Production through Transesterifica-
tion of Triolein with Various Alcohols in an Ultrasonic
Field,” Renewable Energy, Vol. 34, No. 3, 2009, pp. 766-
768. doi:10.1016/j.renene.2008.04.007
[10] K. G. Georgogianni, M. G. Kontominas, E. Tegou, D.
Avlonitis and V. Gergis, “Biodiesel Production: Reaction
and Process Parameters of Alkali-Catalyzed Transesteri-
fication of Waste Frying Oils,” Energy & Fuels, Vol. 21,
No. 5, 2007, pp. 3023-3027. doi:10.1021/ef070102b