You are on page 1of 8

Bioresource Technology 209 (2016) 343–350

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Economical and green biodiesel production process using river snail


shells-derived heterogeneous catalyst and co-solvent method
Wuttichai Roschat a, Theeranun Siritanon a, Teadkait Kaewpuang a, Boonyawan Yoosuk b, Vinich
Promarak c,⇑
a
School of Chemistry, Institute of Science, Suranaree University of Technology, Muang District, Nakhon Ratchasima 30000, Thailand
b
Renewable Energy Laboratory, National Metal and Materials Technology Center (MTEC), National Science and Technology Development Agency, 114 Thailand Science Park,
Phahonyothin Road, Klong 1, Klong Luang, Pathumthani 12120, Thailand
c
Department of Material Science and Engineering, School of Molecular Science & Engineering, Vidyasirimedhi Institute of Science and Technology, Wangchan, Rayong 21210, Thailand

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 River snail shells-derived CaO catalyst


was synthesized for the first time. River snail shells
 98.5% FAME yield was achieved in
90 min under the use 10% v/v of THF Calcination 800 ° C, 3 h
in methanol.
 The transesterification reaction
mechanism is experimentally
demonstrated. CaO catalysts
 The co-solvent method of THF/
methanol successfully decreases
+
10% v/v THF in methanol,
activation energy of reaction. 65 ° C, 90 min
Palm oil Methanol Biodiesel Glycerol

a r t i c l e i n f o a b s t r a c t

Article history: River snail shells-derived CaO was used as a heterogeneous catalyst to synthesize biodiesel via transes-
Received 2 January 2016 terification of palm oil with methanol. The shell materials were calcined in air at 600–1000 °C for 3 h.
Received in revised form 4 March 2016 Physicochemical properties of the resulting catalysts were characterized by TGA–DTG, XRD, SEM, BET,
Accepted 5 March 2016
XRF, FT-IR and TPD. CaO catalyzed transesterification mechanism of palm oil into biodiesel was verified.
Available online 11 March 2016
The effects of adding a co-solvent on kinetic of the reaction and %FAME yield were investigated. %FAME
yield of 98.5% ± 1.5 was achieved under the optimal conditions of catalyst/oil ratio of 5 wt.%; methanol/oil
Keywords:
molar ratio of 12:1; reaction temperature of 65 °C; 10% v/v of THF in methanol and reaction time of
River snail shells
CaO catalyst
90 min. The results ascertained that river snail shells is a novel raw material for preparation of CaO cat-
Biodiesel alyst and the co-solvent method successfully decreases the reaction time and biodiesel production cost.
Reaction mechanism and kinetics Ó 2016 Elsevier Ltd. All rights reserved.
Co-solvent

1. Introduction (Wang et al., 2015). Biodiesel is composed of long chain fatty acid
alkyl esters typically produced from a reaction between a triglyc-
Biodiesel is known to be an alternative to diesel fuel derived eride in fats or oils and short chain alcohols, mainly methanol
from petroleum source. It can be used on its own or mixed with (Leung et al., 2010).
diesel in any diesel–engine vehicles. Advantages of biodiesel are Generally, biodiesel production via transesterification reaction
biodegradability, non-toxicity, and lower CO2 and sulfur emission requires a catalyst to promote the reaction. Commonly, homoge-
neous catalysts such as KOH and NaOH are widely used because
they give high biodiesel yield under mild reaction condition and
⇑ Corresponding author. Tel.: +66 33 014150; fax: +66 33 014445. short reaction time (Lee et al., 2015). However, the use of these cat-
E-mail address: vinich.p@vistec.ac.th (V. Promarak). alysts has drawbacks as large amount of water is required to wash

http://dx.doi.org/10.1016/j.biortech.2016.03.038
0960-8524/Ó 2016 Elsevier Ltd. All rights reserved.
344 W. Roschat et al. / Bioresource Technology 209 (2016) 343–350

the produced biodiesel to eliminate the catalyst and soaps. Homo- were crushed, sieved and calcined at different temperatures (600–
geneous catalysts are also difficult to be reused and they can cor- 1000 °C) with a heating rate of 10 °C/min for 3 h.
rode reactors (Farooq et al., 2013). Such disadvantages increase The elemental compositions of river snail shells and the
the cost of biodiesel production. On the other hand, heterogeneous obtained catalysts were analyzed by a PHILIPS Magi X diffuse
basic catalysts have many advantages as they are cheap, environ- wavelength X-ray Fluorescence (XRF) spectrophotometer with
mentally friendly and non-toxic. It is possible to reuse and recycle 1 kW Rh Ka radiation. Thermal decomposition of the shells was
heterogeneous basic catalysts without employing water to clean analyzed by a Rigaku TG/DTA 8120 thermal analyzer under air flow
the biodiesel product. Moreover, the reaction usually results in with a heating rate of 10 °C/min. Phases of both river snail shells
pure glycerol as a by-product (Farooq et al., 2013; Reyero et al., and the resulting calcined materials was determined by a PHILIPS
2014). X’Pert-MDP X-ray diffractometer using Cu Ka radiation
Among the heterogeneous catalyst, calcium oxide (CaO) is one (k = 1.5418 Å) at 1400 W, 40 kV and 40 mA. JEOL JSM 6010LV Scan-
of the best heterogeneous catalysts with good catalytic activity ning Microscope was used to investigate the samples. The samples
and low solubility in methanol. It is also cheap, environmentally were also analyzed by Fourier transforms infrared (FT-IR) spec-
friendly, easy to prepare and it can be easily recovered and recy- troscopy using a Perkin–Elmer FT-IR spectroscopy spectrum RXI
cled (Witoon et al., 2014; Chen et al., 2014). Many research articles spectrometer.
have reported the study of natural materials such as egg shells Brunauer Emmett Teller (BET) was used to investigate surface
(Wei et al., 2009) and sea shells (Lee et al., 2015; Nakatani et al., area, mean pore diameter and pore volume based on adsorption
2009) as starting materials for CaO preparation via calcination pro- and desorption isotherm of N2 gas on the Bel-sorp-mini II (Bel-
cess. It is obvious from literatures that different types of shells give Japan). Basic strength and basic site properties of calcined river
CaO with different activity. Moreover, reactions catalyzed by shell- snail shells were evaluated by Hammett indicator method and
derived CaO usually take 3–6 h to finish which is still quite slow temperature programmed desorption method (TPD) on the
comparing to homogeneous catalyzed reactions (Boro et al., Chemisorption Analyzer (Belcat B) with CO2 as the probe molecule
2011). It is the focus of this work to study the preparation and cat- (Chen et al., 2014).
alytic activity of CaO catalysts prepared from river snail shells. A
river snail is one kind of freshwater mussels in the family Vivipari- 2.3. Transesterification reaction
dae commonly found around muddy river area in Thailand. Their
fleshy part is consumed and the shells are always wasted. All transesterification reactions were carried out in a three-neck
Another purpose of the current research is to present a strategy round bottom batch reactor equipped with a condenser and a ther-
to accelerate the catalyzed reactions by a co-solvent method. The mocouple. The mixture of river snail shells-derived CaO catalysts
method has been applied to improve the transesterification pro- and methanol was heated at 65 °C and added to palm oil. The reac-
cess. For example, tetrahydrofuran (THF), acetone, diethyl ether tion conditions were designed as follows: methanol to oil molar
and chlorobenzene were used as a co-solvent in homogeneous cat- ratio of 6:1–18:1, amount of catalyst to oil is 1–7 wt.% and mag-
alyzed biodiesel production (Alhassan et al., 2014; Thanh et al., netic stirring speed of 300 rpm. Both the type and the amount of
2013; Soriano et al., 2009; Luu et al., 2014a,b; Babaki et al., the co-solvent were varied. The ratio of co-solvent to methanol
2015). Some reports also show that the co-solvent method can (%v/v) was set at 5% and 10%. After the reaction completed, the
reduce the reaction time and reaction temperature and improve mixture was filtered to separate the solid catalyst. The catalysts
some fuel properties of the produced biodiesel (Luu et al., 2014a, were collected and reused for the next run.
b; Alhassan et al., 2014). To monitor the reaction, 0.5–1.0 mL of the reaction mixture was
In addition, the transesterification reaction with the prepared collected every 30 min for 10 h for analysis. Excess methanol and
CaO is systematically studied in detail including the condition opti- co-solvent in the sampled mixture were evaporated in an oven
mization, reaction mechanism and the reusability of the catalysts. before the analysis of biodiesel yield in term of the fatty acid
The obtained knowledge is expected to deepen the understanding methyl ester yield (%FAME) was performed. Proton nuclear mag-
of the reaction itself and to be useful in real applications. netic resonance (1H NMR) on a Brüker AscendTM 500 MHz spec-
trometer was employed to evaluate %FAME. Tetramethylsilane
(TMS) and CDCl3 were used as the internal reference and a solvent,
2. Methods respectively (Monteiro et al., 2009; Roschat et al., 2012, 2016). A
representative example of 1H NMR spectra and their interpretation
2.1. Materials are presented in the Supporting Information (Fig. S1).

The palm oil (acid value = 0.30 mg KOH g1) and waste cooking
2.4. Study of reaction kinetics
oil (WCO, acid value = 1.54 mg KOH g1) in this work were pur-
chased from commercial sources in Thai market. The fatty acid pro-
In a typical biodiesel production, 3 mol of alcohol react with
file of the palm oil is presented in previous report (Roschat et al.,
1 mol of triglyceride. Methanol is most often used in this reaction
2016). River snail shells were collected from local restaurants.
because of its suitable physical and chemical properties (Mahesh
Methanol (99.5%), acetone (99%), 1-propanol (99%), 2-propanol
et al., 2015). According to Vujicic et al. (2010) and Birla et al.
(99%), tetrahydrofuran (THF 99.5%), ethanol (99%), and ethylene
(2012), this transesterification can be assumed a single step reac-
glycol (99%) were obtained from Fluka. Hammett indicators
tion. The rate law of the reaction can be presented by Eq. (1):
namely phenolphthalein, indigo carmine, 2,4-dinitroaniline and
4-nitroaniline of AR grade purchased from Aldrich and Fluka were d½TG 0
used in this work. r a ¼ ¼ k  ½TG  ½MeOH3 ð1Þ
dt
0
where ½TG; ½MeOH and k are concentration of triglyceride, concen-
2.2. Catalyst preparation and characterization tration of methanol, and the equilibrium rate constant, respectively.
During the reaction, excess amount of methanol was used to shift
The river snail shells were washed with water several times and the equilibrium toward the product thus methanol concentration
air-dried overnight in an oven at 100 °C. The dried river snail shells was considered constant. Therefore, the reaction behaves as a
W. Roschat et al. / Bioresource Technology 209 (2016) 343–350 345

pseudo-first order reaction (Singh and Fernando, 2007; Roschat o 0.6

Derivertive Weight (%/ C)


et al., 2016) and Eq. (1) can be rearranged to: 100 793 C

o
d½TG 0.5
r a ¼ ¼ k  ½TG ð2Þ 90
dt

Weight (%)
CaCO (s)
0 80
3 0.4
where k is modified rate constant ðk ¼ k  ½MeOH3 Þ. Assuming the
initial triglyceride concentration at initial time (t = 0), ½TG0 is 0.3
1 mol and ½TGt is triglyceride concentration at time t, the integra- 70
tion of Eq. (2) gives Eq. (3): CaO(s)+ CO (g) 0.2
2
60
ln½TGt ¼ k  t ð3Þ
0.1
[TG]t is related to %FAME yield by Eq. (4): 50
0.0
½TGt ¼ 1  xME ð4Þ
40
%FAME
 100 200 300 400 500 600 700 800 900 1000
where xME is 100
. Substituting ½TGt in Eq. (3) gives Eq. (5):
o
Temperature ( C)
 ln ½1  xME  ¼ k  t ð5Þ
Fig. 1. TG/DTA profile of the river snail shells.
Therefore, a plot of ln½1  xME  versus t will be linear and its
slope is equal to k. On the other hand, activation energy ðEa Þ and
the frequency factor ðAÞ can be obtained from a slop and an inter- Table 1, calcining river snail shells at higher than 800 °C signifi-
cept of ln k versus 1=T plot, respectively (Birla et al., 2012; Singh cantly reduced BET surface area, total pore volume, and mean pore
and Fernando, 2007; Zhang et al., 2010; Roschat et al., 2016). diameter of the sample because of the severe sintering of the par-
ticles. CO2-TPD of river snail shells calcined at 800 °C showed
2.5. Study of transesterification mechanism higher total basic sites and basic site density (basic site den-
sity = total basic site/BET surface area) than those of the shells cal-
Mechanism of transesterification reaction was studied on a cined at 900 °C and 1000 °C. These results suggested that BET
reaction between 2 g of CaO catalyst obtained from river snail surface area correlates with the basic site of CaO catalysts. The
shells mixed with 20 mL of methanol at 65 °C for 30 min. In this total basic sites are strongly correlated to the catalytic activity of
case, both the liquid mixture and the solid catalysts were analyzed. catalysts as catalysts with higher basic sites usually give higher
0.5 mL of the reaction mixture was sampled and analyzed with 1H biodiesel yield. It is therefore concluded that the optimum calcina-
NMR and 13C NMR. The solid catalysts were separated from reac- tion temperature to prepare CaO catalysts from river snail shells is
tion mixture by filtration (denoted as wet CaO–methanol sample). 800 °C. Similar results were also obtained in reports of Viriya-
A portion of the wet CaO–methanol sample was dried in vacuum empikul et al. (2012) and Lee et al. (2014).
for 10 min and in an oven at 65 °C for another 20 min to give dried The basic strengths (H_value) of all calcined river snail shells
CaO–methanol sample. Then, both of the resulting solid materials measured by Hammett indicator method were in the range of
(wet CaO–methanol and dried CaO–methanol) were analyzed by 15.0–18.4 as evidenced by their ability to change a color of 2,4-
FT-IR and TG–DTA method. In addition, effects of the mixing dinitroaniline (pKa = 15) but not 4-nitroaniline (pKa = 18.4). In
sequence of the reactants (palm oil and methanol) and CaO cata- contrast, The H_value of uncalcined river snail shells was less than
lysts on the reaction kinetics were also investigated. 8.0–10.0. SEM image of river snail shells-derived CaO catalysts cal-
cined at 800 °C for 3 h (Fig. S3 in the Supporting Information)
3. Results and discussion shows that the sample contained large particles with high porosity.

3.1. Catalyst preparation and characterizations 3.2. Optimization of reaction conditions on the transesterification of
palm oil
As expected, XRF results indicated that river snail shells and the
obtained catalysts were composed of more than 98.5% of Ca. Fig. 1 3.2.1. Effects of calcination temperatures of river snail shells
shows TG/DTA analysis of river snail shells. The small weight loss Fig. 2(a) shows the influence of the calcination temperature on
due to the loss of water molecules and organic compounds was the activity of river snail shells-derived CaO catalysts. Catalytic
observed at 50–400 °C. The major weight change occurred at activities in transesterification of palm oil and WCO of the shells
around 600–800 °C. After 800 °C, weigh of the sample remained calcined at different temperatures were investigated. The results
constant at 53 wt.% which was in agreement with the theoretical showed that %FAME yield significantly increased when calcination
weight change (loss 44 wt.%) in the decomposition of CaCO3 to temperature of the shells increased from 600 to 800 °C. On the
CaO (Wei et al., 2009). other hand, increasing the calcination temperature to 900 and
XRD patterns (Fig. S2 in the Supporting Information) of river 1000 °C resulted in the samples with lower catalytic activity which
snail shells calcined at different temperatures indicated that cal- gave lower %FAME yield under the same reaction condition. These
cining the shells at temperature lower than 800 °C (600 and results were expected as the sample calcined at 600 and 700 °C
700 °C) did not lead to the formation of CaO and the sample con- only contained CaCO3 while calcining river snail shells at 900
tained only CaCO3.The samples calcined at 800, 900 and 1000 °C and 1000 °C resulted in CaO with lower BET surface area and total
for 3 h were consisted of CaO as a major phase with no indication basic site (Lee et al., 2015; Viriya-empikul et al., 2012). Hence, the
of CaCO3 remained. Results from TGA and XRD were in agreement optimum calcination temperature to synthesize CaO catalysts from
as both suggested that a temperature of at least 800 °C was river snail shells was 800 °C.
required to transform CaCO3 to CaO. This conclusion is also in
agreement with those reported by Lee et al. (2015). 3.2.2. Effects of catalyst loading amount
The BET surface area, total pore volume and mean pore diame- The effects of catalyst loading amount on the transesterification
ter directly impact the active sites on the catalysts. As shown in of palm oil were tested by varying the amount of river snail
346 W. Roschat et al. / Bioresource Technology 209 (2016) 343–350

Table 1
Physicochemical properties of river snail shells-derived CaO catalyst calcined at different temperatures for 3 h.

Calcination BET surface area total pore volume mean pore Basic sites (mmol/g) Basic sites density Basic
temperature (°C) (m2 g1) (cm3 g1) diameter (nm) (lmol/m2) strengths
Weak Medium Strong Total
<300 °C 300–550 °C >550 °C
800 3.495 1.872  102 21.425 0.006 4.080 4.790 8.876 393.76 15 < H_ < 18.4
900 2.664 6.611  103 9.925 0.066 2.614 5.378 8.058 330.60 15 < H_ < 18.4
1000 0.678 1.515  103 8.940 0.114 2.926 4.786 7.826 86.63 15 < H_ < 18.4

(a) (b) 100


100 97.56
94.52
84.74
80
80 78.61
72.46

Palm oil 66.04


60
% FAME

60 WOC

% FAME
40 Amount of catalyst
40
Cat. 1%
20 Cat. 3%
20
Cat. 5%
8.34
5.62 Cat. 7 %
2.29 1.32 0
0
600 700 800 900 1000 0 60 120 180 240 300 360 420 480 540 600
Calcination temperature (°C) Reaction time (min)

(c) 100 (d)


100
80
90
60
% FAME

80
% FAME

40
70

20
Methanol/oil 60
9:1 12:1
0 15:1 18:1
50
0 60 120 180 240 300 360 1 2 3 4 5 6 7
Reaction time (min) Reused times

Fig. 2. (a) %FAME yield obtained in transesterification using river snail shells-derived CaO catalysts calcined at designated temperatures for 3 h (reaction conditions:
temperature = 65 °C, methanol/oil molar ratio = 12:1, catalyst amount of 5 wt.% for 180 min. (b) Effects of catalyst loading amount on %FAME yield (reaction conditions:
temperature = 65 °C and methanol/oil molar ratio = 12:1). (c) Effects of methanol to oil molar ratio on %FAME yield (reaction conditions: temperature = 65 °C and catalyst
loading amount of 5 wt.%). (d) Reusability study of river snail shells-derived CaO catalyst (reaction conditions: temperature = 65 °C and catalyst loading amount of 5 wt.%,
methanol/oil molar ratio = 12:1 and time of 180 min).

shells-derived CaO catalysts from 1 wt.% to 7 wt.%. %FAME yield 3.2.3. Effects of methanol to oil ratio
increased with increasing catalyst loading amount from 1 wt.% to Fig. 2(c) illustrates variation of %FAME yield with different
5 wt.% (Fig. 2(b)). In general, more CaO catalysts produce methoxide methanol to oil molar ratios using river snail shells-derived
anion as a nucleophile which would then attack electrophilic car- CaO as a catalyst. Yield of biodiesel increased when methanol
bonyl carbon of triglycerides to give biodiesel. Therefore, increasing to oil molar ratio was increased from 9:1 to 12:1. In general,
amount of catalysts generally enhance biodiesel production (Lee increasing methanol to oil molar ratio shifts the reaction equilib-
et al., 2015; Chen et al., 2014). However, the biodiesel yield did rium toward the formation of biodiesel (Viriya-empikul et al.,
not increase when the catalyst loading amount was further 2012; Obadiah et al., 2012). However, increasing the ratio to
increased to 7 wt.%. Such limitation may be related to the phase 15:1 did not change biodiesel yield and further increase metha-
mixing between palm oil, methanol and solid catalyst as high vis- nol to oil ratio to 18:1 decreased %FAME yield. In this case, glyc-
cosity of slurry limited mass transfer of reactants to catalysts (Liu erol by-product may dissolve in the excessive methanol and
et al., 2007; Mahesh et al., 2015; Wu et al., 2013). Therefore, the thereby inhibit the mixing between methanol with the catalysts
optimum catalyst loading was found to be 5 wt.% in this system. and oil. Nevertheless, the optimum methanol to oil molar ratio
W. Roschat et al. / Bioresource Technology 209 (2016) 343–350 347

in this work was 12:1 which resulted in over 95% FAME yield in anion (Lee et al., 2015; Liu et al., 2007). In the present work, such
180 min at 65 °C. reaction mechanism was carefully studied and verified.

3.3. Reusability of river snail shells-derived CaO catalyst 3.4.1. FT-IR analysis
FT-IR spectra of fresh CaO, dried CaO–methanol, wet CaO–
In order to test catalyst reusability, the used catalysts were methanol and liquid methanol was shown in Fig. S5 and the assign-
examined. As depicted in Fig. 2(d), the catalysts could still give ments of each major IR peaks were listed in Table S1 (in the
over 90% FAME yield after 4 uses. Beyond 4 cycles, the catalytic Supporting Information). The wet CaO–methanol sample displayed
performance of the catalyst decreased dramatically. FT-IR spectra a spectrum similar to liquid methanol because most of methanol
of deactive catalysts after being used for 7 times were investigated molecules overwhelmed on the CaO surface by a physical adsorp-
and shown in Fig. S4 in the Supporting Information. The results tion. In addition, the peak at 3660 cm1 associated with CaOAH
indicated that the catalysts reacted with CO2 and water forming stretching vibration was observed as the hydration process with
CaCO3 and Ca(OH)2 as evidenced by FT-IR band sat 1425 cm1 moisture in air changed CaO to Ca(OH)2 (Esipovicha et al., 2014).
(C@O stretching) and 3646 cm1 (CaOAH stretching) (Chen et al., The dried CaO–methanol sample showed bands at 1050–
2014; Roschat et al., 2016). In addition, IR spectra also indicated 1100 cm1, 2700–2900 cm1 and 1350–1500 cm1 associated with
that the reaction mixture including palm oil, methanol, biodiesel CAO stretching, CAH stretching and CH3 bending, respectively.
and glycerol overwhelmed the surface of the catalysts as stretching These peaks significantly shifted comparing to those of liquid
band of OAH at 3378 cm1, CAH at 2927 and 2846 cm1, C@O at methanol. In contrast, OAH stretching band at 3000–3600 cm1
1425 cm1, CAOAC at 1132 cm1, CAOH at 1066 cm1 and C@O was much weaker. From FT-IR results, it can be confirmed that
bending at 872 cm1 are observed (Wei et al., 2009; Roschat basic sites on CaO surface transforms methanol to the more reac-
et al., 2016). tive CH3O and H+ species as explained in Scheme S1 (Fan et al.,
Furthermore, the leaching of CaO catalysts is another cause of 2013; Esipovicha et al., 2014). These species electrostatically bind
catalyst deactivation. In this case, atomic absorption spectroscopy on the surface of the CaO catalyst.
(AAS) analysis was applied to determine the leached calcium (Ca)
ions in biodiesel product as shown in Fig. 3. The leaching was high-
3.4.2. TG/DTA analysis
est in the product obtained from fresh catalysts and gradually
Fig. 4 shows TG/DTA curves of fresh CaO and dried CaO–metha-
decreased with number of cycles. Such decrease in Ca ion leaching
nol (wet CaO–methanol was showed in Fig. S6 in the Supporting
is reasonable as the catalyst surface was covered by the reaction
Information). The results show that weight of the fresh CaO
mixture. Both the Ca leaching and the covered catalyst surface
catalysts remained constant at 99.95% from 35 to 900 °C. Dried
were the main reasons of the catalyst deactivation.
CaO–methanol sample displayed 6% weight loss in the range of
180–280 °C which can be assigned to the loss of CH3O and H+ species
3.4. Study of transesterification mechanism binding on the CaO surface. This confirms that interaction of CH3O
and H+ species with CaO surface was strong. DTA curve shows
Transesterification mechanism is very important in designing endothermic peak maximum at 200 °C and 250 °C corresponding
the reactor, controlling reaction and improving the biodiesel pro- to the leaving of those species, respectively. In addition, TG/DTA
duction process. Several articles reported the possible transesteri- curves of the dried CaO–methanol sample also shows peak in range
fication reaction mechanism via a typical nucleophilic substitution of 350–470 °C which corresponds to the thermal decomposition of
reaction (Zhang et al., 2010; Fan et al., 2013). In the basic catalyzed Ca(OH)2 (Roschat et al., 2016; Yoosuk et al., 2010). This indicates
conditions, the reaction starts by removing a proton (H+) from the that during the activation of the methanol by CaO catalysts some
methanol making methoxide anion (CH3O) as a high reactive parts of the catalytic surface of CaO also undergoes hydrolysis to
nucleophilic specie. The nucleophilic CH3O attacks electrophilic the inactive Ca(OH)2. Thus, TG/DTA data strongly supports the pos-
carbonyl carbon in triglyceride to give biodiesel product and glyc- sibility of CaO catalyzed transesterification mechanism of palm oil
erol by-product (Scheme S1 in the Supporting Information). Hence, and methanol via the formation of methoxide anion on the surface
the key step in this mechanism is the generation of methoxide CaO catalyst surface, and agrees with the FT-IR data.

135
121.7
120 100 0.14
110.2
Derivertive Weight (%/ C)

105 0.12
A
Calcium content (ppm)

90 95 0.10
Weight mass (%)

75 0.08
65.3
60 53.9 52.9 90 0.06
45 B
0.04
30 85
18.9 0.02
15 13.1

0.00
0
80
1 2 3 4 5 6 7
100 200 300 400 500 600 700 800 900
Reused times o
Temperature ( C)
Fig. 3. The leaching of calcium into each of biodiesel product obtained from the
reuse of catalyst. Fig. 4. TG/DTA curves of fresh CaO (A) and dried CaO–methanol (B).
348 W. Roschat et al. / Bioresource Technology 209 (2016) 343–350

3.4.3. NMR analysis into methanol can improve biodiesel production as the mixing
Fig. S7(a) in the Supporting Information shows the 1H NMR between methanol and palm oil will be improved (Soriano et al.,
spectra of liquid methanol (spectrum 1) and dried CaO–methanol 2009; Alhassan et al., 2014; Luu et al., 2014a,b; Dianursanti and
in CDCl3 solvent (spectrum 2). As can be seen from Fig. S7(a), Wijanarko, 2015).
OAH proton of methanol was strongly deshielded when methanol Types and properties of co-solvent used in this study namely ace-
reacted with CaO catalyst because its electron density was reduced tone, 1-propanol, 2-propanol, THF, ethanol and ethylene glycol are
by the two neighboring oxygen atom. On the other hand, high elec- shown in Table S2 (in the Supporting Information) (Green and
tron density on oxygen atom of CH3O specie shielded the CH3 pro- Perry, 2004). According to the obtained results shown in Table 2, %
tons and shifted their peak upfield from chemical shift 3.46 ppm to FAME yield was increased when co-solvents were used in the order:
3.35 ppm. Similarly, such high electron on CH3O species also THF > 1-propanol > 2-propanol > acetone > ethanol > ethylene gly-
shield the carbon atom of the methoxide (CH3O) resulting in an col. The results suggested that physicochemical properties espe-
upfield shift of the 13C NMR peak in dried CaO–methanol as illus- cially polarity and boiling point of THF were suitable to increase
trated in Fig. S7(b). The chemical shift of the methoxide carbon the dissolution of palm oil in methanol. Increased polarity of
(CH3O) shifted from 50.11 ppm to 49.98 ppm. These results con- co-solvents led to a decreased biodiesel yield as the high polar
firm that methanol molecules are transformed into the reactive co-solvents can dissolve well in methanol but not in palm oil. These
methoxide spices during the activation process by CaO catalyst. results agree with the report of Mohammed-Dabo et al. (2012) who
used THF as a co-solvent in transesterification of Jatropha curcas seed
3.4.4. Effect of mixing sequence on the reaction kinetics and %FAME oil catalyzed by sodium hydroxide.
yield As shown in Fig. 6(a), addition of 10% v/v of THF improved %
According to Scheme S1, the first step in the mechanism is FAME yield to 95.90% at the reaction time 90 min while non
methoxide anion formation. To further emphasize the importance
of this step, the effects of mixing sequence of the reaction mixture
Table 2
on %FAME yield were investigated. When methanol was mixed
The effect of co-solvent on the transesterification reaction of palm oil and methanol.
with CaO catalyst before adding palm oil, 94.34% FAME yield was
achieved after 180 min. Such %FAME yield was significantly higher Type of co- %FAME k constant**
solvent (v/v*) (102 min1)
than that obtained from a reaction prepared by mixing CaO with
palm oil (66.67%) before adding methanol (Fig. 5(a)). This differ- Reaction time Reaction time
90 min 120 min
ence clearly shows the significance of methoxide formation. When
CaO catalysts were first mixed with palm oil, their surface includ- Non co-solvent 63.93 80.01 1.836
system
ing active sites was cover by the oil hindering methoxide formation
THF 5% 93.53 97.27 4.002
once methanol was added and consequently slowed the transester- THF 10% 95.03 96.68 4.087
ification reaction. Acetone 5% 70.17 86.89 1.978
The effect of mixing sequence is also reflected in the rate con- Acetone 10% 90.23 95.28 2.765
stants (k) obtained from ln½1  xME  versus reaction time plot as 2-Propanol 5% 81.63 95.81 2.320
2-Propanol 10% 88.44 94.59 2.183
shown in Fig. 5(b). When methanol was first mixed with CaO cat-
1-Propanol 5% 93.46 94.79 2.159
alysts before adding palm oil, k constant was 1.28  102 min1. 1-Propanol 10% 94.44 95.69 2.067
While mixing palm oil with CaO catalysts before adding methanol Ethanol 5% 62.50 89.69 1.960
resulted in a k constant of only 8.26  103 min1. Ethanol 10% 75.47 93.45 2.09
Ethylene glycol 83.33 84.39 1.310
5%
3.5. Effect of co-solvent on transesterification process Ethylene glycol 80.60 85.84 1.340
10%
Phase separation between hydrophilic methanol, hydrophobic *
%v/v relative to amount of methanol.
oil and solid catalyst is generally known to be the major problem **
Reaction conditions: methanol/oil molar ratio of 12:1, catalyst loading amount
in heterogeneous catalyzed reaction. Addition of another solvent of 5 wt.% and 65 °C.

(a) (b)
100 5
MeOH + CaO catalyst
80 4 Oil + CaO catalyst
-ln (1 - XME)

3
60
% FAME

2
40
1
20 MeOH + CaO catalyst
Oil + CaO catalyst 0
0
0 60 120 180 240 300 360 0 60 120 180 240 300 360
Reaction time (min) Reaction time (min)
Fig. 5. (a) The effect of mixing sequence of reaction mixture on %FAME yield and (b) kinetics study (reaction condition: methanol/oil molar ratio of 12:1, catalyst loading
amount of 5 wt.% and temperature of 65 °C).
W. Roschat et al. / Bioresource Technology 209 (2016) 343–350 349

(a) (b)
-3.0
100 Non co-solvent
-3.5 Slope = -8.13
80 -4.0 Interept = 20.71

-4.5
60
% FAME

lnk
-5.0
40 -5.5

-6.0
20 THF co-solvent
Non co-solvent Slope = -6.95
-6.5
THF co-solvent Interept = 16.69
0 -7.0
0 30 60 90 120 150 180 2.9 3.0 3.1 3.2 3.3
3 -1
Reaction time (min) 1/T x 10 (K )

Fig. 6. (a) %FAME of with and without 10% v/v of THF systems. (b) Arrhenius plots of ln k versus 1/T for transesterification of palm oil with and without 10% v/v THF systems
(reaction conditions: methanol/oil molar ratio of 12:1, catalyst loading amount of 5 wt.% and 65 °C).

co-solvent system gives only 63.49% FAME yield. The obtained k Babaki, M., Yousefi, M., Habibi, Z., Mohammadi, M., Brask, J., 2015. Effect of water,
organic solvent and adsorbent contents on production of biodiesel fuel from
value of 10% v/v THF co-solvent system was 4.09  102 min1
canola oil catalyzed by various lipases immobilized on epoxy-functionalized
which was twice larger than that of non co-solvent system silica as low cost biocatalyst. J. Mol. Catal. B: Enzym. 120, 93–99.
(1.84  102 min1). The effects of THF on reaction kinetics were Birla, A., Singh, B., Upadhyay, S.N., Sharma, Y.C., 2012. Kinetics studies of synthesis
also observed in the decrease of activation energy (57.79 kJ/mol) of biodiesel from waste frying oil using a heterogeneous catalyst derived from
snail shell. Bioresour. Technol. 106, 95–100.
and increase in frequency factor (1.17  107 min1) comparing to Boro, J., Thakur, A.J., Deka, D., 2011. Solid oxide derived from waste shells of
those of non co-solvent reaction (67.60 kJ/mol and 9.87  108 min1) Turbonilla striatula as a renewable catalyst for biodiesel production. Fuel
as shown in Fig. 6(b). Process. Technol. 92, 2061–2067.
Chen, G., Shan, R., Shi, J., Yan, B., 2014. Ultrasonic-assisted production of biodiesel
from transesterification of palm oil over ostrich eggshell-derived CaO catalysts.
Bioresour. Technol. 171, 428–432.
4. Conclusion
Dianursanti, Religia P., Wijanarko, A., 2015. Utilization of n-hexane as co-solvent to
increase biodiesel yield on direct transesterification reaction from marine
Transesterification of palm oil to biodiesel product was cat- microalgae. Procedia Environ. Sci. 23, 412–420.
alyzed by river snail shells-derived CaO catalysts. The reaction time Esipovicha, A., Danova, S., Belousova, A., Rogozhin, A., 2014. Improving methods of
CaO transesterification activity. J. Mol. Catal. A: Chem. 395, 225–233.
of transesterification was decreased from 180 min in non co- Fan, M.M., Huang, J.L., Yang, J., Zhang, P.B., 2013. Biodiesel production by
solvent reaction to 90 min in a co-solvent reaction with 10% v/v transesterification catalyzed be an efficient choline ionic liquid catalyst. Appl.
THF. Additionally, the possible transesterification mechanism was Energy 108, 333–339.
Farooq, M., Ramli, A., Subbarao, D., 2013. Biodiesel production from waste oil using
experimentally investigated. The results indicate that river snail bifunctional heterogeneous solid catalysts. J. Cleaner Prod. 59, 131–140.
shells-derived CaO catalysts and THF co-solvent are both capable Green, D.W., Perry, R.H., 2004. Perry’s Chemical Engineering Handbook, Seventh ed.
of improving transesterification of palm oil in biodiesel production. Mcgraw-Hill, pp. 732–2293.
Lee, H.V., Juan, J.C., Abdullah, N.F.B., Taufiq-Yap, Y.H., 2014. Heterogeneous base
catalysts for edible palm and non-edible Jatropha-based biodiesel production.
Acknowledgements Chem. Cent. J. 8, 1–9.
Lee, S.L., Wong, Y.C., Tan, Y.P., Yew, S.Y., 2015. Transesterification of palm oil to
biodiesel by using waste obtuse horn shell-derived CaO catalyst. Energy
This work was financially supported from Thailand Graduate Convers. Manage. 93, 282–288.
Institute of Science and Technology (TGIST 01-55-011), National Leung, D.Y.C., Wu, X., Leung, M.K.H., 2010. A review on biodiesel production using
Science and Technology Development Agency (NSTDA) – Thailand catalyzed transesterification. Appl. Energy 87, 1083–1095.
Liu, X.J., He, H.Y., Wang, Y.J., Zhu, S.L., 2007. Transesterification of soybean oil to
and Suranaree University of Technology. Thanks for the support biodiesel using SrO as a solid base catalyst. Catal. Commun. 8, 1107–1111.
from the Thailand Research Fund (IUG5180010) – Thailand and Luu, P.D., Takenaka, N., Luu, B.V., Pham, L.N., Imamura, K., Maeda, Y., 2014a. Co-
the Biofuel Development for Thailand funded through Center for solvent method produce biodiesel from waste cooking oil with small pilot plant.
Energy Procedia 61, 2822–2832.
Innovation in Chemistry (PERCH-CIC), PERDO, CHE, Thailand. Luu, P.D., Troung, H.T., Luu, B.V., Pham, L.N., Imamura, K., Takenaka, N., Maeda, Y.,
2014b. Production of biodiesel from Vietnamese Jatrophacuecas oil by a co-
solvent method. Bioresour. Technol. 173, 309–316.
Appendix A. Supplementary data Mahesh, S.E., Ramanathan, A., Meera, K.M.S.B., Narayanan, A., 2015. Biodiesel
production from waste cooking oil using KBr impregnated CaO as catalyst.
Supplementary data associated with this article can be found, in Energy Convers. Manage. 91, 442–450.
Mohammed-Dabo, I.A., Ahmad, M.S., Hamza, A., Muazu, K., Aliyu, A., 2012. Co-
the online version, at http://dx.doi.org/10.1016/j.biortech.2016.03. solvent transesterification of Jatropha curcas seed oil. J. Pet. Technol. Altern.
038. Fuels 3 (4), 42–51.
Monteiro, M.R., Ambrozin, A.R.P., Liao, L.M., Ferreira, A.G., 2009. Determination of
biodiesel blend levels in different diesel samples by 1H NMR. Fuel 88, 691–696.
References Nakatani, N., Takamori, H., Takeda, K., Sakugawa, H., 2009. Transesterification of
soybean oil using combusted oyster shell waste as a catalyst. Bioresour.
Alhassan, Y., Kumar, N., Bugaje, I.M., Pali, H.S., Kathkar, P., 2014. Co-solvents Technol. 100, 1510–1513.
transesterification of cotton seed oil into biodiesel: effects of reaction Obadiah, A., Swaroopa, G.A., Kumar, S.V., Jeganathan, K.R., Ramasubbu, A., 2012.
conditions on quality of fatty acids methyl esters. Energy Convers. Manage. Biodiesel production from palm oil using calcined waste animal bone as
84, 640–648. catalyst. Bioresour. Technol. 116, 512–516.
350 W. Roschat et al. / Bioresource Technology 209 (2016) 343–350

Reyero, I., Arzamendi, G., Gandía, L.M., 2014. Heterogenization of the biodiesel Wang, Y.T., Fang, Z., Zhang, F., Xue, B.J., 2015. One-step production of biodiesel from
synthesis catalysis: CaO and novel calcium compounds as transesterification oils with high acid value by activated Mg–Al hydrotalcite nanoparticles.
catalysts. Chem. Eng. Res. Des. 9 (2), 1519–1530. Bioresour. Technol. 193, 84–89.
Roschat, W., Kacha, M., Yoosuk, B., Sudyoadsuk, T., Promarak, V., 2012. Biodiesel Wei, Z., Xu, C., Li, B., 2009. Application of waste eggshell as low-cost solid catalyst
production based on heterogeneous process catalyzed by solid waste coral for biodiesel production. Bioresour. Technol. 100, 2883–2885.
fragment. Fuel 98, 194–202. Witoon, T., Bumrungsalee, S., Vathavanichkul, P., Palitsakun, S., Saisriyoot, M.,
Roschat, W., Siritanon, T., Yoosuk, B., Promarak, V., 2016. Biodiesel production from Faungnawakij, K., 2014. Biodiesel production from transesterification of palm
palm oil using hydrated lime-derived CaO as a low-cost basic heterogeneous oil with methanol over CaO supported on bimodal meso–macroporous silica
catalyst. Energy Convers. Manage. 108, 459–467. catalyst. Bioresour. Technol. 156, 329–334.
Singh, A.K., Fernando, S.D., 2007. Reaction kinetics of soybean oil transesterification Wu, H., Zhang, J., Wei, Q., Zhenge, J., Zhang, J., 2013. Transesterification of oil to
using heterogeneous metal oxide catalysts. Chem. Eng. Technol. 30, 1716–1720. biodiesel using zeolite supported CaO as strong base catalysts. Fuel Process.
Soriano, U.N., Venditti, R., Argyropoulos, S.D., 2009. Biodiesel synthesis via Technol. 109, 13–18.
homogenous Lewis-acid-catalyzed transesterification. Fuel 88, 560–565. Yoosuk, B., Udomsap, P., Puttasawat, B., Krasae, P., 2010. Modification of calcite by
Thanh, T.L., Okitsu, K., Sadanga, Y., Takenaka, N., Maeda, Y., Bandow, H., 2013. A new hydration–dehydration method for heterogeneous biodiesel production
co-solvent method for the green production of biodiesel fuel-optimization and process: the effects of water on properties and activity. Chem. Eng. J. 162,
practical application. Fuel 103, 742–748. 135–141.
Viriya-empikul, N., Krasae, P., Nualpaeng, W., Yoosuk, B., Faungnawakij, K., 2012. Zhang, L., Sheng, B., Xin, Z., Liu, Q., Sun, S., 2010. Kinetics of transesterification of
Biodiesel production over Ca-based solid catalysts derived from industrial palm oil and dimethyl carbonate for biodiesel production at the catalysis of
wastes. Fuel 92, 239–244. heterogeneous base catalyst. Bioresour. Technol. 101, 8144–8150.
Vujicic, Dj., Comic, D., Zarubica, A., Micic, R., Boskovi, G., 2010. Kinetics of biodiesel
synthesis from sunflower oil over CaO heterogeneous catalyst. Fuel 89, 2054–
2061.

You might also like