Anda di halaman 1dari 10

Energy Conversion and Management 119 (2016) 453–462

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Rice husk-derived sodium silicate as a highly efficient and low-cost basic


heterogeneous catalyst for biodiesel production
Wuttichai Roschat a,⇑, Theeranun Siritanon 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

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

Article history: In the present work, rice husk-derived sodium silicate was prepared and employed as a solid catalyst for
Received 24 February 2016 simple conversion of oils to biodiesel via the transesterification reaction. The catalyst was characterized
Received in revised form 19 April 2016 by TG–DTA, XRD, XRF, FT-IR, SEM, BET and Hammett indicator method. Under the optimal reaction con-
Accepted 20 April 2016
ditions of catalyst loading amount of 2.5 wt.%, methanol/oil molar ratio of 12:1, the prepared catalysts
gave 97% FAME yield in 30 min at 65 °C, and 94% FAME yield in 150 min at room temperature. The trans-
esterification was proved to be pseudo-first order reaction with the activation energy (Ea) and the fre-
Keywords:
quency factor (A) of 48.30 kJ/mol and 2.775  106 min1 respectively. Purification with a cation-
Sodium silicate
Rice husk
exchange resin efficiently removed all soluble ions providing high-quality biodiesel product that meets
Heterogeneous catalyst all the ASTM and EN standard specifications. Rice husk-derived sodium silicate showed high potential
Transesterification to be used as a low-cost, easy to prepare and high performance solid catalyst for biodiesel synthesis.
Biodiesel Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction On the other hand, heterogeneous catalysts can be easily sepa-


rated through filtration, reused and recycled several times and
As issues relating to energy and environment are of main con- require no neutralization or washing process. In addition they
cern in the past decades, research on alternative and sustainable can also produce a high purity glycerol by-product. Hence,
energy has gained considerable interest. Biodiesel (fatty acid heterogeneous-catalyst is an efficient and more cost-effective tool
methyl ester, FAME) is a good candidate for replacing petrol–diesel to produce biodiesel [7,10,11]. There are several reports on biodie-
fuel because it can be used in any diesel-engine without modifica- sel production using heterogeneous catalysts such as Mo–Mn/c-
tion [1–3]. It is also clean, renewable, biodegradable, environmen- Al2O3–MgO [12], KOH/bentonite [13], Li4SiO4 [14], LiAlO2 [15]
tally friendly and inexpensive [4,5]. Generally, biodiesel can be and CaO–La2O3 [16]. However, some of these catalysts still have
produced via a direct transesterification reaction of vegetable oils to be developed because they show low catalytic activity, compli-
or animal fats and alcohols in the presence of a catalyst [6–8]. cate to generate and absorb CO2 and H2O in air easily.
Commonly, KOH, NaOH and CH3ONa are widely used as a homoge- Calcium oxide (CaO) is one of the most studied catalysts mate-
neous basic catalyst because they have high catalytic activity rial as it has lower price, non-toxicity, less solubility in methanol
which can complete the reaction in 1 h under mild conditions at and high catalytic activity. They can be prepared from natural
40–60 °C [9]. Nevertheless, the use of these catalysts still involves waste obtuse horn [3], waste coral fragment [2], hydrated lime
many problems such as difficult catalyst separation, soap forma- [17], ostrich eggshells [18,19], waste oyster shells [20], chicken
tion and reactor corrosion. In addition, large amount of water is bones [21], waste shells of egg [18,22], golden apple snail shells
usually required to wash the biodiesel product to eliminate the and meretrix venus shells [22]. Although CaO catalysts offer many
catalysts which leads to the increase in the overall production cost advantages, in some case, the catalyzed reactions still require long
and the environmental problems [3]. reaction times (at least 3–8 h) which increased the production cost.
Sodium silicate (Na2SiO3) with formula Na2OnSiO2 is a well-
known water–glass or liquid glass which can be easily prepared
from SiO2 and NaOH [23,24]. Na2SiO3 was also used as a starting
⇑ Corresponding authors.
reactant to synthesize mesoporous silica material [25], n-alkane/
E-mail addresses: roschat1@gmail.com (W. Roschat), vinich.p@vistec.ac.th
(V. Promarak). silica composite phase [26] and b-zeolite [27]. Recently, several

http://dx.doi.org/10.1016/j.enconman.2016.04.071
0196-8904/Ó 2016 Elsevier Ltd. All rights reserved.
454 W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462

articles report the use of calcined sodium silicate as a solid basic


catalyst in the transesterification of vegetable oil (soybean oil,
rapeseed oil and jatropha oil) with methanol. Sodium silicate is
an effective catalyst for transesterification under mild reaction
condition and short reaction time (60 min at 60 °C). Furthermore,
the catalyst can be reused for at least 5 times without loss of activ-
ity [28–31].
An important focus of this work is to present a simplified way to
synthesize sodium silicate from rice husk which is scrap from agri-
cultural and to use the obtained material as a low-cost basic
heterogeneous catalyst in biodiesel production. The rice husk-
derived sodium silicate were characterized and tested in the pro-
duction of biodiesel using oils and methanol. Several parameters,
such as catalyst amount, methanol/oil molar ratio, reaction tem-
perature, type of oil and free fatty acid (FFAs) quantity and
reusability were investigated. Fuel properties of the obtained bio-
diesel after purification and treatment process were evaluated by
American Society for Testing and Material (ASTM) methods and
European Standard methods (EN14214) for bio-auto fuels. In addi-
tion, reaction kinetics of transesterification using sodium silicate
obtained from rice husk was investigated and compared with Fig. 1. Flow diagram of the sodium silicate preparation.
CaO catalysts.

(SEM) was performed on JEOL JSM 5410LV scanning electron


2. Materials and methods microscope at an accelerated voltage of 20 kV. Brunauer Emmett
Teller (BET) was employed on a Bel-sorp-mini II (Bel-Japan) to
2.1. Materials investigate surface area and pore volume. The samples were char-
acterized by Fourier transform infrared (FT-IR) spectroscopy with
Many kinds of oils containing various amounts of FFAs includ- Perkin-Elmer FT-IR spectroscopy spectrum RXI spectrometer in
ing palm oil (acid value of 0.32 mg KOH/g oil), lard oil (acid value the range of 450–4000 cm1 with resolution of 4 cm1 and potas-
of 0.42 mg KOH/g oil), coconut oil (acid value of 0.28 mg KOH/g sium bromide (KBr) was used as a matrix. Hammett indicator
oil), sunflower oil (acid value of 0.20 mg KOH/g oil), soybean oil method was applied to test the basic strength by the following
(acid value of 0.27 mg KOH/g oil), rice bran oil (acid value of indicators with different acidity functions (H_); phenolphthaleine
0.18 mg KOH/g oil), jatropha oil (acid value of 1.31 mg KOH/g oil) (H_ = 9.8), indigo carmine (H_ = 12.2), 2,4-dinitroaniline
and waste cooking oil (acid value of 2.31 mg KOH/g oil) were (H_ = 15.0) and 4-nitroaniline (H_ = 18.4). The amount of basicity
obtained from commercial sources in Thailand and were used was measured by benzoic acid titration method using various
without any purification. The analytical grade methanol, Hammett Hammett indicators [28].
indicators (phenolphthalein, thymolphthalein, indigo carmine, 2,4-
dinitroaniline and 4-nitroaniline) and sodium hydroxide (NaOH) 2.3. Catalytic tests and product analysis
were purchased from Fluka and Sigma–Aldrich Chemical. Calcium
oxide (analytical grade CaO_AR) was purchased from Acros Chem- The transesterification was carried out in a three-neck round
ical Co. Ltd. Rice husk collected from a local rice mill was washed bottom batch reactor equipped with a reflux condenser, a magnetic
with deionized water and dried at 110 °C overnight before uses. stirrer (300 rpm) and a thermocouple. Palm oil was used as starting
material to optimize reaction condition. A mixture of solid catalyst
2.2. Catalyst preparation and characterization and methanol were preheated at designated temperature (55, 60,
65 and 70 °C) and added to 30 mL of oil. The transesterification
Fig. 1 shows a flow diagram of the sodium silicate preparation. reaction was performed with various catalysts loading amount
The dried rice husk was first digested by reflux method using 1 M (0–5.0 wt.% relative to oil weight) and methanol to oil molar ratios
HCl at 100 °C for 3 h, washed with water several times and dried at (3:1–11:1).
110 °C overnight. After that, white rice husk ash (RHA) was To follow the reaction progress, 0.5 mL of solution mixture was
obtained by calcining the digested rice husk at 700 °C for 3 h to sampled and heated in an oven at 80 °C for 3 h to remove excess
remove organic contains as shown in Fig. S1 in the Supporting methanol before biodiesel yield analysis. Proton nuclear magnetic
Information [32]. The carbon residue content in the obtained resonance (1H NMR) on a Brüker AscendTM 500 MHz spectrometer
RHA was determined by a CHN 628 LECO Instruments. Next, 10 g was employed to evaluate biodiesel yield in term of the fatty acid
of the obtained RHA and 100 mL of 1 M NaOH were mixed and methyl ester yield (%FAME). Tetramethylsilane (TMS) and CDCl3
boiled in a covered 250 mL Erlenmeyer flask for 1 h with constant were used as the internal reference and a solvent, respectively
stirring. After that, the resulting solution was dried in an oven at [2,17]. The example of the obtained 1H NMR spectrum is presented
110 °C for 6 h. Finally, the resulting materials were calcined in a in Fig. S2 in the Supporting Information. The biodiesel yield was
furnace at designated temperature (200–500 °C) in air for 1 h to calculated in term of %FAME according Eq. (1):
generate sodium silicate materials. 2ACH3
Both the uncalcined and calcined materials were analyzed by X- %FAME ¼  100 ð1Þ
3ACH2
ray powder diffraction (XRD) using a Bruker D5005 X-ray diffrac-
tometer with Cu Ka radiation (k = 1.5418 Å). Elemental composi- where ACH3 is an integration of the methoxy protons of the methyl
tions of the samples were analyzed by a PHILIPS Magi X ester moiety (CH3AOCOA) at 3.66 ppm chemical shift. ACH2 is an
wavelength dispersive X-ray Fluorescence (XRF) spectrophotome- integration of the proton in a-carbonyl methylene groups (RACH2-
ter with 1 kW Rh Ka radiation. Scanning electron microscopy AOCOA) both in triglyceride and methyl ester at chemical shift of
W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462 455

2.30 ppm. The factors 3 and 2 were derived from the number of perature resulted in a slightly lower surface area. Such difference is
attached protons at the methoxy and a-carbonyl methylene car- to be expected because at high temperature atoms in the small
bons, respectively [2,17]. particles could diffuse to the boundary and fuse the particles
In addition, the final biodiesel product was determined accord- together forming larger particles. Similar results were observed
ing to EN14214 standard method using a gas chromatograph (GC- in report of Lee et al. [3]. Although, all sodium silicate samples
2010, Shimadzu) and the example of the obtained GC chro- obtained from rice husk displayed low surface area, they show high
matogram is illustrated in Fig. S3 in the Supporting Information. basic strength as measured by Hammett indicator method which
The amount of sodium ion leached from the catalyst in the demonstrated H_ values in the range of 15.0–18.4 as evidenced
obtained biodiesel was analyzed by atomic absorption spectropho- by its coloration with 2,4-dinitroaniline (pKa = 15) but not with
tometry (AAS) [2,33]. 4-nitroaniline (H_ = 18.4). The total basicity of sodium silicate cat-
The obtained biodiesel product was filtered to separate catalyst, alyst obtained from rice husk was the order of calcined sample at
heated to remove excess methanol and treated with commercially 300 °C > 400 °C > 200 °C > 500 °C, respectively (Table 1). The strong
available cation-exchange resins (DowexÒ 50WX8) to eliminate basic strength and high basicity of the catalysts are correlated to
sodium ion [17]. The treated biodiesel were tested by American the catalytic activity of catalysts as catalysts with higher basic
Society for Testing and Material (ASTM D6751) methods and Euro- strength and basicity usually give higher biodiesel yield [4,28]. Ele-
pean Standard methods (EN14214) for bio-auto fuels [17,34,35]. mental analysis of all calcined samples with XRF spectroscopy indi-
cated that they were composed of mainly SiO2 and Na2O with
approximately 1% of other elements as shown in Table 1.
3. Results and discussion
3.2. Optimization of reaction conditions on the transesterification of
3.1. Catalyst characterization palm oil

Thermogravimetric analysis was used to determine the opti- 3.2.1. Effects of calcining temperature of catalysts
mum calcination temperature of the digested rice husk to prepare The effects of calcining temperature of rice husk-derived
RHA (Fig. S5 in the Supporting Information). As seen from Fig. S4, sodium silicate on the catalytic activity were evaluated. As shown
the weight loss occurred in the range of 30–400 °C which indicated in Fig. 4(a), FAME yield was only 84.39% when the sample was cal-
the removal of moisture and organic compounds such as cellulose, cined at 200 °C. This may be due to an incomplete conversion of
hemicelluloses and lignin. Based on TGA results, temperature at the sample to sodium silicate crystalline phase as indicated by
700 °C was chosen for calcining the digested rice husk to produce the low-intensity peaks in XRD pattern (Fig. 1(b)). XRD patterns
RHA [32]. The obtained RHA show low carbon content of 0.65 wt. of the sample calcined at 300, 400 and 500 °C exhibit clear and
%. The result indicated that the acid and thermal treatment of rice sharp peaks, indicating more crystallinity. As a consequence, bio-
husk can be produced the high purity of RHA. diesel yield was significantly increased to 98.04% when calcining
After boiled and dried, the mixture of RHA and NaOH was ana- temperature of the sample was increased from 200 to 300 °C. How-
lyzed by TG–DTA as shown in Fig. 2(a). There was only one major ever, a further increase in calcining temperature resulted in a
weight loss at below 200 °C which correspond to water molecules slightly decreased %FAME yield under the same reaction condition.
removal. Therefore, the temperature of 200, 300, 400, and 500 °C This reduction of %FAME yield was caused by the sintering effect
were selected to optimize the calcining temperature. As depicted during calcination which resulted in the lowering of surface area
in Fig. 2(b), the XRD pattern of RHA shows amorphous phase with (Table 1). As a result, the optimal calcining temperature to synthe-
a broad curve. On the other hand, mixture of RHA and NaOH trans- size sodium silicate catalysts from rice husk was 300 °C.
formed to sodium silicate after calcination as their XRD patterns
exhibit the characteristic peaks of Na2SiO3 crystalline phase (PDF 3.2.2. Effects of catalyst amount
00-016-0818). Similar results were also reported by Guo et al. [28]. The effects of catalyst loading amount on conversion of palm oil
The samples were analyzed by FT-IR spectroscopy as depicted to biodiesel were studied and displayed in Fig. 4(b). In transester-
in Fig. 2(c). Spectra of all the calcined samples showed strong broad ification mechanism, the basic sites of sodium silicate generate a
absorption bands at 999 cm1 and 884 cm1 corresponding to the reactive nucleophile (methoxide anion) from methanol which will
stretching vibrations of SiAO and SiAOASi, respectively. In addi- react with electrophilic carbonyl carbon of triglyceride to produce
tion, the absorption peaks at 510 cm1, 710 cm1 and 1465 cm1 biodiesel. Therefore, as the amount of catalyst increases, active
were assigned to SiAO bending, SiAOASi bending and Si@O basic site will generally increase and the biodiesel product is
stretching vibrations, respectively. These vibrations indicated the enhanced [3,21]. As a result, the biodiesel yield increased from
presence of SiAOASi linking structure thus confirmed the forma- 49.37% to 97.80% when the catalyst content was increased from
tion of sodium silicate [28,30]. In addition, broad peaks at 3000– 0.5 wt.% to 2.5 wt.%. However, beyond 2.5 wt.% of catalyst loading
3600 cm1 were attributed to SiAOAH stretching vibrations amount, the biodiesel yield did not increase due to the limitation
caused by the absorbed water molecules on the surface. The sur- of mass transfer and the high viscosity of the reaction mixture,
face morphology of sodium silicate calcined at 300 °C was studied both of which, lead to the problem of mixing each reactant compo-
by SEM as depicted in Fig. 2(d) which illustrates agglomerated and nent [15,16].
large particles size of few microns. The formation of methoxide species is very important in biodie-
N2 adsorption–desorption isotherm of all sodium silicate sam- sel production. In case of heterogeneous catalyzed reaction, the
ples derived from rice husk are displayed in Fig. 3(a)–(d). The methoxide species are formed on the surface of the catalysts. It
adsorption–desorption of all samples is a typical type II pattern is, therefore, very important that methanol is absorbed on the cat-
which is a characteristic of a nonporous or macroporous material alyst surface. This process depends on mass transfer of the metha-
based on IUPAC’s classification [36]. The large particles of about nol hence depends on how well the mixture which includes solid
few microns as seen from SEM image, the low BET surface area catalyst, liquid methanol, and oil are mixed [37,38]. In fact, it
and the low pore volume (Table 1) are all consistent with this clas- was found that methanol adsorption is the rate-determining step
sification as well. [39]. Loading the solid catalysts into the reaction mixture increases
The BET surface area of all samples was very similar. Neverthe- viscosity of the mixture making it more difficult to get proper
less, the slight difference can be observed as higher calcining tem- mixed. At one point, the advantage of having more catalysts in
456 W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462

(a) 1.2 (b)


0

Derivative Weight (%/ C)


-5 TGA curve 1.0

o
o
DTA curve 500 C

Intensity (a.u.)
-10 0.8
Weight loss %

o
400 C
-15 0.6
o
300 C
-20 0.4
o
200 C
-25 0.2
RHA
Na2SiO3
-30 0.0
50 100 150 200 250 300 350 400 450 500 20 25 30 35 40 45 50 55 60 65 70
o
Tempature ( C) 2 theta (degree)

Si-O-Si
Si-O
(c) Si-O-H
Si-O
Si-O-Si
710
510
(d)
999
3440 884

Si=O
o 1465
500 C
Transmission (a.u.)

o
400 C

o
300 C

o
200 C

RHA Si=O Si-O-Si


1555 1104

4000 3000 2000 1000


-1
Wavenumber (cm )

Fig. 2. (a) TG/DTA curves of the resulting materials obtained from mixing and boiling the mixture of RHA and NaOH solution. (b) XRD patterns and (c) IR spectra of RHA and
calcined samples. (d) SEM image of the sodium silicate obtained from calcining the mixture of RHA and NaOH solution at 300 °C for 1 h.

the reaction will be outweighed by the drawbacks of poor mixing. slightly decreased biodiesel yield as the evaporated methanol
Therefore, the catalyst loading amount should be optimized and in inhibited the reaction on the three-phase interfaces namely liquid
this work, 2.5 wt.% catalyst loading was used to further optimize phase (palm oil), gas phase (methanol) and solid phase (catalysts)
transesterification condition. [18,42].

3.2.3. Effects of methanol to oil molar ratio 3.3. Effects of oil type
The effects of methanol to oil molar ratio on the biodiesel yield
were investigated. As illustrated in Fig. 4(c), the %FAME yield grad- Free fatty acid (FFAs) value (mg KOH/g oil) in the oil has a sig-
ually increased as the molar ratio of methanol to palm oil increased nificant effect on the biodiesel yield during transesterification
from 6:1 to 12:1. Further increase in the methanol to oil molar reaction because high FFAs content causes soap formation which
ratio beyond 12:1, slightly decreased the conversion of palm oil reduces catalyst activity [29,40,43]. Many kinds of oils containing
to biodiesel. At high methanol to oil molar ratio, the glycerol by- various amounts of FFAs namely palm oil, rice bran oil, coconut
product would dissolve in the excessive methanol and impede oil, lard oil, soybean oil, sunflower oil, jatropha oil and waste cook-
the reaction of methanol with oil and catalyst. Additionally, the ing oil were used to investigate the effects of FFAs on the biodiesel
polar hydroxyl group in methanol behaves as emulsifier which yield as seen in Fig. 5. The results showed that type of oils did not
makes it more difficult to separate biodiesel product from the mix- have any effects on the FAME yield in this case as they all contain
ture resulting in the reduced %FAME yield [28,40,41]. In this case, only small amount of FFAs content. In all cases, the FAME yields
the optimum methanol to oil molar ratio was, therefore, 12:1, was maintained at over 93% even for jatropha and waste cooking
which was 4 times more than the required stoichiometric molar oil which have the highest FFAs.
ratio (3:1).
3.4. Reusability of the catalysts
3.2.4. Effects of reaction temperature
Reaction temperature is one of the most important parameters The reusability (without cleaning process) of rice husk-derived
which directly influence both the reaction rate and yield. In this sodium silicate catalyst was investigated using the optimized con-
work, effects of reaction temperature were examined by varying ditions. As illustrated in Fig. 6, %FAME yield was still higher than
from 45 °C to 75 °C with constant reaction time (30 min). As pre- 93% when the catalysts were reused for four times and the value
sented in Fig. 4(d), %FAME yield was increased from 42.32% to decreased to 83% after the fifth time. The possible reason for the
96.15% when reaction temperature increased from 45 °C to 65 °C. loss of catalytic activity was the leaching of active species
Further increase the temperature to over 65 °C which is above [44,45]. In this case, the AAS analysis was used to evaluate the lea-
the boiling point of methanol (64.7 °C at 1 atm) resulted in the ched sodium (Na) ions (Fig. 6). It is obvious from Fig. 6 that Na ion
W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462 457

(a) (b)
0 0
28 200 C 28 300 C

Volume adsorped (cm /g STP )

Volume adsorped (cm /g STP )


24 24

3
20 20

16 16

12 12

8 8

4 4

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
P/P0 P/P0

(c) (d)
0 0
28 400 C 28 500 C
Volume adsorped (cm /g STP )

Volume adsorped (cm /g STP )


24 24
3

3
20 20

16 16

12 12

8 8

4 4

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
P/P0 P/P0

Fig. 3. N2 adsorption/desorption plots of sodium silicate derived from rice husk with different calcining temperatures of (a) 200 °C, (b) 300 °C, (c) 400 °C and (d) 500 °C.

Table 1
Physicochemical properties of sodium silicate derived from rice husk calcined at different temperatures.

Sample BET surface area (m2/g) Pore volume (cm3/g) Total basicity (mmol/g) Element composition (%)
SiO2 Na2O Al2O3 MgO Other
Calcined at 200 °C 1.31 2.08 10.620 68.98 30.07 0.21 0.11 0.63
Calcined at 300 °C 1.14 1.39 11.298 68.96 30.10 0.22 0.12 0.60
Calcined at 400 °C 1.05 1.65 10.864 69.04 29.97 0.19 0.09 0.71
Calcined at 500 °C 0.95 1.30 9.962 69.05 30.01 0.19 0.09 0.66

was leached to the reaction mixture. The leaching decreased with the reaction mixture namely palm oil, biodiesel, glycerol and
reusing times but continue to occur. Although the catalysts methanol may cover the surface of the catalysts [3,17].
remained active with some leaching in the first four runs, the activ- Although leaching of Na ions from the catalysts seem to be the
ity decreased after five times. It is believed that the catalysts were major cause of efficiency reduction of the catalysts in the transes-
able to function with some degree of Na loss but when Na contin- terification reaction, the catalysts could be regenerated just by a
ued to leach, the catalysts will eventually deactivated. reaction with NaOH. Guo et al. [30] presented that the first, second
Fig. 7(a) compares XRD profiles of fresh catalysts and catalysts and third time regenerated catalysts still performed excellently to
that have been used for 9 times. The sharpness of diffraction peaks give over 96.5% biodiesel yield. Hence, the possibility of using rice
is lower in the used catalysts indicating loss of crystalline phases husk-derived sodium silicate as a high potential, low-cost and
during the reaction possibly due to Na leaching. In addition, FT- renewable catalysts are high.
IR spectra in Fig. 7(b) support the structure change of sodium sili-
cate by showing absorption bands at 710 cm1 and 884 cm1 with 3.5. Catalytic activity comparison of rice husk-derived sodium silicate,
lower intensity, relating to lower number of SiAOASi bonds. In CaO and NaOH
contrast, absorption bands at 999 cm1 and 1465 cm1 which cor-
respond to SiAO bending and Si@O stretching vibrations have Fig. 8 shows comparison of catalytic performance of rice husk-
increased intensity, confirming that some SiAOASi bonds were derived sodium silicate, CaO analytical grade (AR), CaO derived
broken down during the reusing processes of the catalysts from eggshell (heterogeneous catalyst) and NaOH (homogeneous
[28,30]. Furthermore, the broader absorption band at 3440 cm1 catalyst) under optimized reaction conditions. The results showed
(SiAOAH stretching) in the used catalysts indicates that sodium that FAME yield reached 96% within 15 min when NaOH catalyst
silicate catalyst reacted with water molecules in the reactants was used, while the reaction catalyzed by rice husk-derived
(palm oil and methanol) which is to be expected as the reaction sodium silicate showed FAME yield of 97% after 30 min. For the
was carried out in an open air. In addition, liquid compounds in reaction catalyzed by CaO both analytical grade (AR) and derived
458 W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462

(a) 100 98.04 (b) 100


95.24
88.76
84.39
80 80

% FAME
% FAME 60 60

40 40

20 20

0 0
200 300 400 500 0 1 2 3 4 5
Calcination temperature (°C) Catalyst loading (wt%)

(c) (d)
100 96.15 100
92.17 93.02

80 77.33 80

% FAME
60 60
% FAME

40 40

20 20

0 0
6 9 12 15 45 50 55 60 65 70 75
Methanol/oil molar ratio Temperature (°C)

Fig. 4. Effect of reaction conditions on %FAME yield. (a) Calcination temperature of catalyst (catalyst loading 2 wt.%, methanol/oil molar ratio of 9:1, 60 °C and reaction time of
30 min). (b) Catalyst loading amount (calcined the catalyst at 300 °C, methanol/oil molar ratio of 9:1, 60 °C and reaction time of 30 min). (c) Methanol to oil molar ratio
(calcined the catalyst at 300 °C, catalyst loading 2.5 wt.%, 60 °C and reaction time of 30 min). (d) Reaction temperature (calcined the catalyst at 300 °C, catalyst loading 2.5 wt.
%, methanol/oil molar ratio of 12:1 and reaction time of 30 min).

100
100 300

80 Type of oil Acid value


250
(mg KOH/g oil) 80

Sodium content (ppm)


rice bran 0.18
60
% FAME

sunflower 0.20 200


soybean 0.27
% FAME

60
coconut 0.28
40 palm 0.32 150
lard 0.42
waste cooking 2.31 40
20 jatropha 1.31 100

20 FAME
0 50
Sodium content
ja kig
su ran

lm

as lard
so er

t
co n

ha
nu
ea
w

op
o
pa
co
b

yb

co
lo

tr
ce

nf

0 0
te
ri

1 2 3 4 5 6 7 8 9
w

Type of oil Reused times


Fig. 5. Effects of FFAs contents on %FAME yield (calcined the catalyst at 300 °C, Fig. 6. Effects of reuse of the catalyst on the %FAME and analysis amount of Na ion
catalyst loading 2.5 wt.%, methanol/oil molar ratio of 12:1, 65 °C and 30 min). leaching content in biodiesel product (calcined the catalyst at 300 °C, catalyst
loading 2.5 wt.%, methanol/oil molar ratio of 12:1, 65 °C and reaction time of
30 min).
from eggshell, FAME yield reached about 96% after 180 min. In this
research, it was found that the catalytic performance for conver-
sion of palm oil to biodiesel was in the order of NaOH > sodium sil- In both sodium silicate and CaO case, the same active specie,
icate > CaO. It is generally known that homogeneous basic catalysts CH3O, is generated but the mechanism of the generation is differ-
especially NaOH and KOH exhibit higher catalytic activity than ent. Sodium silicate generates CH3O by exchanging its Na+ ion
heterogeneous basic catalyst [40,46,47]. Although both are hetero- with H+ of methanol while O2 on CaO surface extracts H+ from
geneous catalyst, the catalytic activity of sodium silicate was sig- H2O to form OH which then extracts H+ from methanol to form
nificantly higher than that of CaO. Similar results were also CH3O. The difference in mechanism is most likely the reason of
reported by Guo et al. [28]. the different activity. In addition, Guo et al. [30] also reported that
W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462 459

(a) (b)
Si-O-Si Si-O
710 510
Si-O-Si
884
fresh

Transmission (a.u.)
Si=O

Intensity (a.u)
1465
Si-O
999
fresh Si-O-H
3440

th
after 9 reuse
th
after 9 reuse
20 30 40 50 60 70 80 4000 3000 2000 1000
-1
2 theta (degree) Wavenumber (cm )

Fig. 7. (a) XRD patterns and (b) IR spectra of rice husk-derived sodium silicate catalyst after 9th reuse compared with fresh the catalyst.

100 100

80 80

60 60
% FAME

% FAME
40 40
NaOH
rice husk-derived Na 2SiO3
20 20
CaO_AR NaOH
CaO_eggshell rice husk-derived Na2SiO3
0 0
0 30 60 90 120 150 180 210 240 0 30 60 90 120 150 180 210 240
reaction time (min) reaction time (min)

Fig. 8. Comparison of catalytic efficiency of rice husk-derived sodium silicate Fig. 9. Comparison of catalytic performance between rice husk-derived sodium
against CaO and NaOH. Reaction condition of rice husk-derived sodium silicate: silicate and NaOH under room temperature conditions. Reaction condition of rice
calcined the catalyst at 300 °C, catalyst loading 2.5 wt.%, methanol/oil molar ratio of husk-derived sodium silicate: calcined the catalyst at 300 °C, catalyst loading
12:1 and 65 °C. Reaction condition of CaO_AR and CaO_eggshell as calcined the 2.5 wt.%, methanol/oil molar ratio of 12:1. Reaction condition of NaOH: catalyst
catalyst at 800 °C for 3 h [22]: catalyst loading 6 wt.%, methanol/oil molar ratio of loading 0.75 wt.% and methanol/oil molar ratio of 6:1 [47].
15:1 and 65 °C [17]. Reaction condition of NaOH: catalyst loading 0.75 wt.%,
methanol/oil molar ratio of 6:1 and 60 °C [45].
plify the kinetic modeling, the reaction could be assumed pseudo-
first order and the reaction rate can be explained by Eq. (2). In this
CaO may easily be packed by the glycerol by-product which inhib- case, excess methanol was used to shift the reaction equilibrium
ited CH3O generation by interrupting the contact between metha- towards the product thus the reverse reduction is insignificant
nol and basic sites of CaO. and will not be taken into account [17,50,51].
In addition, the room-temperature transesterification reaction
of palm oil catalyzed by rice husk-derived sodium silicate and
NaOH catalysts were investigated as shown in Fig. 9. It was clear 5
that biodiesel yield reached 96% after 90 min (1.5 h) for reaction rice husk-derived Na 2SiO3
catalyzed by NaOH, while the reaction catalyzed by rice husk-
4 CaO_AR
derived sodium silicate gave 94% FAME yield after 150 min
(2.5 h). Reducing production temperature and time means reduc-
ing the production cost. Generally, the use of heterogeneous cata- 3
-ln (1-X ME)

lysts such CaO [2,8,22], KOH/bentonite [13], LiAlO2 [15], CaO-


based/Au [42], natural dolomites [48] and MgOALi2O [49] to pro- 2
duce biodiesel were performed at higher temperature (60–65 °C)
and longer reaction time (2–6 h). Hence, it is also more cost-
effective to apply rice husk-derived sodium silicate as a catalyst 1
for biodiesel production at room temperature.
0
0 10 20 30 40 50 60
3.6. Kinetics of transesterification reaction catalyzed by rice husk-
derived sodium silicate Time (min)

Fig. 10. Comparison of kinetics study between using rice hush-derived sodium
To determine the kinetics of transesterification reaction, the silicate and CaO_AR catalyst under the same reaction condition (reaction condition:
effects of reaction temperature and time were investigated. To sim- catalyst loading 2.5 wt.%, methanol/oil molar ratio of 12:1 and 65 °C).
460 W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462

(a) 5 (b)
o
30 C -2.5
o
4 40 C
o
50 C
o -3.0
-ln (1-X ME) 60 C
3

ln k
-3.5
2
Slope = -5.80897
-4.0 Interept = 14.8362
1
-4.5 Ea = 48.30 KJ/mol
A = 2.775 x 10 min
6 -1
0
0 10 20 30 40 50 60 3.0 3.1 3.2 3.3
3 -1
Time (min) 1/T x 10 (K )

Fig. 11. (a) ln½1  xME  versus reaction time plot at different temperatures; reaction condition: catalyst loading amount of 2.5 wt.%, methanol/oil molar ratio of 12:1. (b)
Arrhenius plot ln k versus 1/T  103 (K1) for transesterification of palm oil using rice husk-derived sodium silicate.

Table 2
Fuel properties of biodiesel obtained from transesterification reaction under optimum condition catalyzed with rice husk-derived sodium silicate after purification and treatment
process.

Fuel properties Unit Standard Palm oil biodiesel


Methyl ester content % 96.5 98.6
Density @ 15 °C kg/m3 860–900 871
Kinematic viscosity@ 40 °C cSt 3.5–5.0 4.15
Flash point °C >120 183
Acid value mg KOH/g oil 60.5 0.19
Water content %w/w oil 60.050 0.012
Oxidation number h >6 16.16
Total contamination mg/kg (ppm) 624 18.9
Copper strip corrosion – Number 1 Number 1
Cloud point °C (3.0)12.0 5.0
Pour point °C (15.0)10.0 2.0
Sulfated ash %w/w oil 60.02 0.0137
Carbon residue %w/w oil 60.05 0.0381
Sodium content ppm 65 3.957

d½TG the optimized reaction condition. Rate constant of reaction cat-


r¼ ¼ k½TG ð2Þ
dt alyzed by rice husk-derived sodium silicate (7.49  102 min1)
is 6 times higher than CaO_AR (1.28  102 min1).
r is the transesterification reaction rate and k is the rate constant
Fig. 11(a) shows rate constant (k) of reaction catalyzed by rice
(k = k0 [ROH]3). [TG] and t are triglyceride concentration and reaction
husk-derived sodium silicate at various temperatures. The
time, respectively. Integrating and rearranging Eq. (2) from t = 0,
obtained k is then used in Arrhenius plot (Fig. 11(b)) which gives
½TG ¼ ½TG0 till t = t, ½TG0 ¼ ½TGt gives Eq. (3):
Ea and A of the reaction. Activation energy (Ea) obtained in this
ln½TG0  ln½TGt ¼ kt ð3Þ work is smaller than that obtained from CaO catalysts as reported
by Birla et al. (79 kJ/mol) [50], Maneerung et al. (83.9 kJ/mol) [51]
ln½TG0 is initial concentration of triglyceride which is assumed and Kostić et al. (108.8 kJ/mol) [52].
to be 1 M, thus ln½TG0 ¼ 0. While ½TGt is related to %FAME yield by

½TGt ¼ 1  xME where xME is %FAME100
. Substituting ½TGt in Eq. (3)
3.7. Physicochemical properties of the palm oil biodiesel
gives Eq. (4):

ln½1  xME  ¼ k  t ð4Þ The physicochemical properties of the biodiesel obtained from
transesterification of palm oil catalyzed by rice husk-derived
By fitting the experimental data to Eq. (4), the slope is equal to sodium silicate after purification with a cation-exchange resin to
rate constant (k) which depends on reaction temperature. The remove all soluble ions were measured by ASTM and EN14214
Arrhenius plot of lnk versus T1 in Eq. (5) can be used to calculate standard methods for bio-auto fuels and shown in Table 2. The pro-
the activation energy (Ea) and frequency factor (A) for the transes- duced biodiesel has high quality standard properties which are
terification reaction. within the standard limits in terms of methyl ester content, den-
sity, viscosity, flash point, acid value, water content, oxidation
Ea 1
lnk ¼   þ lnA ð5Þ number, total contamination, copper strip corrosion, cloud point,
R T
pour point, sulfated ash carbon residue and ion content. Conse-
R is the gas constant (8.314 J mol1 K1) and T is the reaction tem- quently, all results in this work indicated that the low-cost rice
perature (K) [51]. husk-derived sodium silicate catalysts demonstrate good catalytic
Fig. 10 compares rate constant (k) of reaction using rice husk- performance for biodiesel production and should be further pro-
derived sodium silicate that obtained from CaO_AR catalysts under moted in industrial applications.
W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462 461

4. Conclusion [16] Mahesh SE, Ramanathan A, Begum KMMS, Narayanan A. Biodiesel production
from waste cooking oil using KBr impregnated CaO as catalyst. Energy Convers
Manage 2015;91:442–50.
In summary, an efficient sodium silicate solid basic catalyst was [17] Roschat W, Siritanon T, Yoosuk B, Promarak V. Biodiesel production from palm
synthesized by using rice husk and NaOH solution. The prepared oil using hydrated lime-derived CaO as a low-cost basic heterogeneous
catalyst. Energy Convers Manage 2016;108:459–67.
catalysts were used in transesterification reaction of many kinds
[18] Tan YH, Abdullah MO, Hipolito CN, Taufiq-Yap YH. Waste ostrich- and chicken-
of oils to biodiesel. Under optimized reaction conditions, FAME eggshells as heterogeneous base catalyst for biodiesel production from used
yield reached 97% after only 30 min at 65 °C and 94% after cooking oil: catalyst characterization and biodiesel yield performance. Appl
Energy 2015;60:58–70.
150 min (2.5 h) at room temperature. This sodium silicate can be
[19] Chen G, Rui S, Jiafu S, Yan B. Ultrasonic-assisted production of biodiesel from
reused for four times with excellent activity without regeneration transesterification of palm oil over ostrich eggshell-derived CaO catalysts.
or treatment. The obtained biodiesel product after purification and Bioresour Technol 2014;171:428–32.
treatment process by cation-exchange resin shows high quality [20] Nakatani N, Takamori H, Takeda K, Sakugawa H. Transesterification of soybean
oil using combusted oyster shell waste as a catalyst. Bioresour Technol
fuel properties according to ASTM D6751 and EN 14214 standards 2009;100:1510–3.
indicating that the low-cost rice husk-derived sodium silicate cat- [21] Farooq M, Ramli A, Naeem A. Biodiesel production from low FFA waste cooking
alysts has high catalytic performance and could be used for biodie- oil using heterogeneous catalyst derived from chicken bones. Renew Energy
2015;76:362–8.
sel production in industrial scale. [22] Viriya-empikul N, Krasae P, Nualpaeng W, Yoosuk B, Faungnawakij K. Biodiesel
production over Ca-based solid catalysts derived from industrial wastes. Fuel
Acknowledgements 2012;92:239–44.
[23] Halasz I, Agarwal M, Li R, Miller N. Monitoring the structure of water soluble
silicates. Catal Today 2007;126:196–202.
The authors acknowledge the scholarship support from Thai- [24] Arantes RS, Lima RMF. Influence of sodium silicate modulus on iron ore
land Graduate Institute of Science and Technology (TGIST 01-55- flotation with sodium oleate. Int J Miner Process 2013;125:157–60.
[25] Yun-yu Z, Xiao-xuan L, Zheng-xing C. Rapid synthesis of well-ordered
011), National Science and Technology Development Agency mesoporous silica from sodium silicate. Powder Technol 2012;226:239–45.
(NSTDA) – Thailand and Suranaree University of Technology – [26] He F, Wang X, Wu D. Phase-change characteristics and thermal performance of
Thailand. The work is financially supported by Thailand Research form-stable n-alkanes/silica composite phase change materials fabricated by
sodium silicate precursor. Renew Energy 2015;74:689–98.
Fund (IUG5180010) – Thailand.
[27] Selvam T, Bandarapu B, Mabande GTP, Toufar H, Schwieger H. Hydrothermal
transformation of a layered sodium silicate, kanemite, into zeolite Beta (BEAT).
Appendix A. Supplementary material Microporous Mesoporous Mater 2003;64:41–50.
[28] Guo F, Peng ZG, Dai JY, Xiu ZL. Calcined sodium silicate as solid base catalyst
for biodiesel production. Fuel Process Technol 2010;91:322–8.
Supplementary data associated with this article can be found, in [29] Long YD, Guo F, Fang Z, Tian XF, Jiang LQ, Zhang F. Production of biodiesel and
the online version, at http://dx.doi.org/10.1016/j.enconman.2016. lactic acid from rapeseed oil using sodium silicate as catalyst. Bioresour
Technol 2011;102:6884–6.
04.071. [30] Guo F, Wei NN, Xiu ZL, Fang Z. Transesterification mechanism of
soybean oil to biodiesel catalyzed by calcined sodium silicate. Fuel
References 2012;93:468–72.
[31] Kouzu M, Yamanaka S, Hidaka J, Tsunomori M. Heterogeneous catalysis of
calcium oxide used for transesterification of soybean oil with refluxing
[1] Onoji ES, Iyuke SE, Igbafe AI, Nkazi DB. Rubber seed oil: a potential renewable
methanol. Appl Catal: A 2009;355:94–9.
source of biodiesel for sustainable development in sub-Saharan Africa. Energy
[32] Kongmanklang C, Rangsriwatananon K. Hydrothermal synthesis of high
Convers Manage 2016;110:125–34.
crystalline silicalite from rice husk ash. J Spectrosc 2015;2015. Hindawi
[2] Roschat W, Kacha M, Yoosuk B, Sudyoadsuk T, Promarak V. Biodiesel
Publishing Corporation, 5 pages 696513.
production based on heterogeneous process catalyzed by solid waste coral
[33] Long YD, Fang Z, Su TC, Yang Q. Co-production of biodiesel and hydrogen from
fragment. Fuel 2012;98:194–202.
rapeseed and Jatropha oils with sodium silicate and Ni catalysts. Appl Energy
[3] Lee SL, Wong YC, Tan YP, Yew SY. Transesterification of palm oil to biodiesel by
2014;113:1819–25.
using waste obtuse horn shell-derived CaO catalyst. Energy Convers Manage
[34] Mathimani T, Uma L, Prabaharan D. Homogeneous acid catalyzed
2015;93:282–8.
transesterification of marine microalga Chlorella sp. BDUG 91771 lipid – an
[4] Huang R, Cheng J, Qiu Y, Li T, Zhou J, Cen K. Using renewable ethanol and
efficient biodiesel yield and its characterization. Renew Energy
isopropanol for lipid transesterification in wet microalgae cells to produce
2015;81:523–33.
biodiesel with low crystallization temperature. Energy Convers Manage
[35] Abedin MJ, Kalam MA, Masjuki HH, Sabri MFM, Rahman SMA, Sanjid A, et al.
2015;105:791–7.
Production of biodiesel from a non-edible source and study of its combustion,
[5] Chen SY, Lao-ubol S, Mochizuki T, Abe Y, Toba M, Yoshimura Y. Transformation
and emission characteristics: a comparative study with B5. Renew Energy
of non-edible vegetable oils into biodiesel fuels catalyzed by unconventional
2016;88:20–9.
sulfonic acid-functionalized SBA-15. Appl Catal: B 2014;485:28–39.
[36] Khemthong P, Luadthong C, Nualpaeng W, Changsuwan P, Tongprem P, Viriya-
[6] Chen SY, Lao-ubol S, Mochizuki T, Abe Y, Toba M, Yoshimura Y. Production of
empikul N, et al. Industrial eggshell wastes as the heterogeneous catalysts for
Jatropha biodiesel fuel over sulfonic acid-based solid acids. Bioresour Technol
microwave-assisted biodiesel production. Catal Today 2012;190:112–6.
2014;157:346–50.
[37] Xie W, Peng H, Chen L. Transesterification of soybean oil catalyzed by
[7] Xie W, Zhao L. Heterogeneous CaO–MoO3–SBA-15 catalysts for biodiesel
potassium loaded on alumina as a solid-base catalyst. Appl Catal A
production from soybean oil. Energy Convers Manage 2014;79:34–42.
2006;300:67–74.
[8] Sirisomboonchai S, Abuduwayiti M, Guan G, Samart C, Abliz S, Hao X, et al.
[38] Arzamendi G, Campoa I, Arguinarena E, Sánchez M, Montes M, Gandia LM.
Biodiesel production from waste cooking oil using calcined scallop shell as
Synthesis of biodiesel with heterogeneous NaOH/alumina catalysts:
catalyst. Energy Convers Manage 2015;95:242–7.
comparison with homogeneous NaOH. Chem Eng J 2007;134:123–30.
[9] Tubino M, Junior JGR, Bauerfeldt GF. Biodiesel synthesis: a study of the
[39] Dossin TF, Reyniers MF, Marin GB. Kinetics of heterogeneously MgO-catalyzed
triglyceride methanolysis reaction with alkaline catalysts. Catal Commun
transesterification. Appl Catal B 2006;61:35–45.
2016;75:6–12.
[40] Chen GY, Shan R, Shi JF, Yan BB. Transesterification of palm oil to
[10] Taufiq-Yap YH, Teo SW, Rashid U, Islam A, Hussien MZ, Lee KT.
biodiesel using rice husk ash-based catalysts. Fuel Process Technol 2015;
Transesterification of Jatropha curcas crude oil to biodiesel on calcium
133:8–13.
lanthanum mixed oxide catalyst: effect of stoichiometric composition.
[41] Wang JX, Chen KT, Wu JS, Wang PH, Huang ST, Chen CC. Production of
Energy Convers Manage 2014;88:1290–6.
biodiesel through transesterification of soybean oil using lithiumorthosilicate
[11] Sanchez M, Marchetti JM, Boulifi NE, Aracil J, Martinez M. Kinetics of Jojoba oil
solid catalyst. Fuel Process Technol 2012;104:167–73.
methanolysis using a waste from fish industry as catalyst. Chem Eng J
[42] Moushoul EB, Farhadi K, Mansourpanah Y, Nikbakht AM, Molaei R, Forough M.
2015;262:640–7.
Application of CaO-based/Au nano particles as heterogeneous nano catalysts
[12] Farooq M, Ramli A, Subbarao D. Biodiesel production from waste oil using
in biodiesel production. Fuel 2016;164:119–27.
bifunctional heterogeneous solid catalysts. J Clean Prod 2013;59:131–40.
[43] Fadhil AB, Aziz AM, Al-Tamer MH. Biodiesel production from Silybum
[13] Soetaredjo FE, Ayucitra A, Ismadji S, Maukar AL. KOH/bentonite catalysts for
marianum L. seed oil with high FFA content using sulfonated carbon catalyst
transesterification of palm oil to biodiesel. Appl Clay Sci 2011;53:341–6.
for esterification and base catalyst for transesterification. Energy Convers
[14] Dai YM, Chen KT, Chen CC. Study of the microwave lipid extraction from
Manage 2016;108:255–65.
microalgae for biodiesel production. Chem Eng J 2014;250:267–73.
[44] Shan R, Chen G, Yan B, Shi J, Liu C. Porous CaO-based catalyst derived from PSS-
[15] Dai YM, Wu JS, Chen CC, Chen KT. Evaluating the optimum operating
induced mineralization for biodiesel production enhancement. Energy Convers
parameters on transesterification reaction for biodiesel production over a
Manage 2015;106:405–13.
LiAlO2 catalyst. Chem Eng J 2015;280:370–6.
462 W. Roschat et al. / Energy Conversion and Management 119 (2016) 453–462

[45] Syazwani ON, Rashid U, Yap YHT. Low-cost solid catalyst derived from waste [49] Lu H, Yu X, Yang S, Yang H, Tu ST. MgO–Li2O catalysts templated by a PDMS–
Cyrtopleura costata (angel wing shell) for biodiesel production using PEO comb-like copolymer for transesterification of vegetable oil to biodiesel.
microalgae oil. Energy Convers Manage 2015;101:749–56. Fuel 2016;165:215–23.
[46] Luu PD, Troung HT, Luu BV, Pham LN, Imamura K, Takenaka N, et al. Production [50] Birla A, Singh B, Upadhyay SN, Sharma YC. Kinatics studies of synthesis of
of biodiesel from Vietnamese Jatropha cuecas oil by a co-solvent method. biodiesel from waste frying oil using a heterogeneous catalyst derived from
Bioresour Technol 2014;73:309–16. snail shell. Bioresour Technol 2012;106:95–100.
[47] Alhassan Y, Kumar N, Bugaje IM, Pali HS, Kathkar P. Co-solvents [51] Maneerung T, Kawi S, Wang CH. Biomass gasification bottom ash as a source of
transesterification of cotton seed oil into biodiesel: effects of reaction CaO catalyst for biodiesel production via transesterification of palm oil. Energy
conditions on quality of fatty acids methyl esters. Energy Convers Manage Convers Manage 2015;92:234–43.
2014;84:640–8. [52] Kostić MD, Bazargan A, Stamenković OS, Veljković VB, McKay G. Optimization
[48] Nur SZA, Taufiq-Yap YH, Nizah RMF, Teo SH, Syazwani ON, Islam A. Production and kinetics of sunflower oil methanolysis catalyzed by calcium oxide-based
of biodiesel from palm oil using modified Malaysian natural dolomites. Energy catalyst derived from palm kernel shell biochar. Fuel 2016;163:304–13.
Convers Manage 2014;78:738–44.

Anda mungkin juga menyukai