You are on page 1of 9

Energy Conversion and Management 122 (2016) 535543

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Seashell-derived mixed compounds of Ca, Zn and Al as active and stable


catalysts for the transesterification of palm oil with methanol to
biodiesel
Wayu Jindapon a, Siyada Jaiyen b, Chawalit Ngamcharussrivichai b,c,
a
Program in Petrochemistry and Polymer Science, Faculty of Science, Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand
b
Fuels Research Center, Department of Chemical Technology, Faculty of Science, Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand
c
Center of Excellence on Petrochemical and Materials Technology (PETROMAT), Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand

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

Article history: Heterogeneous base catalysts, composed of mixed calcium, zinc and aluminium compounds (ZSA), for the
Received 6 April 2016 transesterification of palm oil with methanol to produce biodiesel as fatty acid methyl esters (FAME)
Received in revised form 4 June 2016 were prepared from waste mixed seashells in the presence of zinc nitrate and alumina via dissolution-
Accepted 6 June 2016
precipitation method. The effects of the catalyst calcination temperature on the physicochemical and
Available online 14 June 2016
catalytic properties were studied. Calcination of ZSA at 300 C generated calcium hydroxide as the main
active phase, while the formation of CaO at higher calcination temperatures contributed to the catalyst
Keywords:
basicity and catalytic activity. The catalyst calcined at 500 C (ZSA-500) possessed a calcium-rich surface
Seashells
Calcium oxide
on which nanocrystallites of CaO were dispersed. The ZSA-500 catalyst showed the highest stability to
Transesterification ambient air (relative humidity of 73%) exposure for 3 d. The loss of catalytic activity at the higher calci-
Biodiesel nation temperatures of 700 and 900 C was related to the formation of Ca12Al14O33 and the transforma-
tion of CaO to calcium carbonate during air exposure. The ZSA-500 catalyst had the highest surface area
and basicity and gave the highest FAME yield (98.0 1.60 wt%). Suitable conditions for transesterification
over ZSA-500 at 60 C were a catalyst loading of 10 wt%, methanol:oil molar ratio of 30:1 and reaction
time of 3 h. Moreover, the ZSA-500 catalyst could be reused after methanol washing to give a FAME yield
more than 96% for at least five uses.
2016 Elsevier Ltd. All rights reserved.

1. Introduction experiences soap formation, catalyst exhaustion, yield loss and


crude glycerol contamination [2].
Due to environmental concerns, the substitution of conven- Heterogeneous catalysis has attracted considerable attention
tional diesel with biodiesel has been increasingly promoted world- for biodiesel production by transesterification owing to the expedi-
wide [1]. Besides being a potentially sustainably renewable raw ency of catalyst recovery, catalyst reusability, process simplicity
material, biodiesel based upon fatty acid methyl esters (FAME) and the higher purity of the obtained FAME and crude glycerol
possesses similar fuel properties to petroleum diesel but it is [4]. A large number of solid catalysts have been investigated for
non-toxic and biodegradable. Moreover, biodiesel combustion pro- the transesterification of model triglycerides or vegetable oils with
duces a relatively clean exhaust gas [2,3]. Nevertheless, the pro- methanol [525]. Calcium oxide (CaO) is a promising catalyst
duction of biodiesel via transesterification, the conventional because of its high basicity, fast reaction rate, low toxicity and
route, of the triglycerides in vegetable oil or animal fats with low cost. The sources for CaO catalyst preparation are available
methanol in the presence of a homogeneous base catalyst is prob- in abundance, for example natural calcium rocks, as calcite and
lematic due to a large amount of alkaline wastewater discharged dolomite [5], and waste calcium materials, such as abalone [6],
from the catalyst removal step [3]. In addition, the process duck eggs [7] and angel wing shell [8]. However, CaO is succes-
sively transformed to calcium methoxide and calcium glyceroxides
(Ca(C3H7O3)2) during use [9]. Although the glyceroxides were
believed to be the active phase of the pure CaO catalyst [10], their
Corresponding author at: Department of Chemical Technology, Faculty of formation contributed to homogeneous catalysis [9,10]. It is
Science, Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand.
worth noting that the solubility in methanol of Ca(C3H7O3)2, at
E-mail address: Chawalit.Ng@chula.ac.th (C. Ngamcharussrivichai).

http://dx.doi.org/10.1016/j.enconman.2016.06.012
0196-8904/ 2016 Elsevier Ltd. All rights reserved.
536 W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543

6216 mg L1 [10], is about 200-fold higher than that of CaO, at give a 1:0.6:0.4 mass ratio of the calcined seashells: Zn(NO3)2
28 mg L1 [9]. Besides, a practical use of CaO as commercial biodie- 6H2O:Al2O3. After stirring for another 60 min, the resulting
sel catalyst is limited by a fast deactivation via chemisorption of mixture was sonicated at 40 C for 4 h until it formed a paste.
atmospheric carbon dioxide (CO2) and moisture upon exposure The moisture remaining in the mixture was removed by drying
to ambient air [9,10]. in an oven at 100 C for 24 h. Finally, the dried solid was crushed
Various supports have been applied to immobilize CaO to into powder (<10 lm) and calcined in a muffle furnace at the
improve the dispersion of the CaO crystallites as well as the textu- desired temperature (300900 C) for 2 h at a heating rate of
ral properties and stability of the immobilized catalysts [1114]. 3 C min1. The catalyst attained was designated as ZSA-x, where
Combining CaO with other metal oxides, such as those of magne- x represents the calcination temperature used.
sium (MgO) [15], zinc (ZnO) [16] and zirconium [17], as binary
mixed oxides is an effective route to improve the specific surface 2.3. Catalyst characterization
area, basicity and catalytic performance of the CaO-based catalysts.
Moreover, three-component mixed oxides of Ca, Mg and Zn have Elemental analysis of the waste mixed seashells was performed
been reported to be highly active catalysts in the methanolysis of using a JEOL ED-2000 energy dispersive X-ray fluorescence spec-
palm kernel oil to FAME [18]. A stable CaO catalyst with an trometer. The crystalline phases present in the catalysts were
enhanced mechanical strength was obtained by forming a Ca/Al determined by powder X-ray diffraction (XRD) using a Rigaku
composite of CaO and Ca12Al14O33 at high temperatures [11], DMAX 2200/Ultima + diffractometer equipped with Cu Ka radia-
whilst a spinel mixed ZnO and aluminium oxide (Al2O3) catalyst tion with a 0.02 step size at room temperature and recording
was preferred in the first commercial biodiesel production plant the spectra over a 2h range of 1080. The diffraction peaks were
using heterogeneous catalysis [4]. assigned after consulting the Joint Committee on Powder Diffrac-
In this study, new active solid catalysts with a higher degree of tion Standard (JCPDS) files. The skeletal structure and the surface
structural stability in ambient air and during the transesterification functional groups of the catalysts attained at different calcination
of vegetable oil with methanol were prepared. As an attempt temperatures were identified by Fourier transform-infrared spec-
towards using more sustainable sources for the catalyst prepara- troscopy (FTIR) using a PerkinElmer Spectrum One instrument.
tion, waste mixed seashells from a frozen food industry were used The samples were prepared according to the standard KBr method.
as the raw material. The catalysts, composed of various types of The spectra were recorded at room temperature for 64 scans over
mixed metal compounds of Ca, Zn and Al (ZSA), were prepared 4004000 cm1 at a resolution of 4 cm1. The thermal decomposi-
via dissolution and precipitation under acidic conditions, followed tion behavior of the catalysts (10 mg) was investigated by
by calcination. The effects of the calcination temperature and of thermogravimetric/differential thermal analysis (TG/DTA) on a
exposing the calcined catalysts to ambient air on the crystalline PerkinElmer Diamond thermogravimeter with a temperature ramp
structure and the catalytic activity were investigated. Overall, rate of 8 C min1 from room temperature to 1000 C under a dry
despite their high catalytic activity, the optimal ZSA catalyst was nitrogen (N2) flow rate of 10 mL min1. The morphology and the
relatively stable in the humid ambient with a relative humidity surface elements of the catalysts were analyzed by scanning elec-
(RH) of 73%, regenerable and reusable. tron microscopy (SEM) with energy dispersive X-ray spectrometry
(EDS) using a JEOL JSM-5800LV microscope equipped with an
Oxford Link/ISIS 300 energy dispersive X-ray spectrometer.
2. Experimental
The total basicity of the catalyst was determined by pulse
chemisorption of CO2 using a Micromeritics AutoChem II 2920
2.1. Materials and reagents
chemisorption analyzer. The catalyst pretreatment was performed
at 300500 C, depending on the calcination temperature of the
Waste mixed seashells of Perna viridis, Amusium pleuronectes,
catalyst, for 2 h under an argon (Ar) flow (50 mL min1). Subse-
Anadara granosa and Meretrix meretrix were provided by the Thai
quently, a small amount of CO2 (10 vol% in Ar) was injected into
Dolomite Co. Ltd., (Surat Thani Province, Thailand). Elemental anal-
the sample as the adsorption step. The excess CO2 was then
ysis by X-ray fluorescence spectroscopy indicated the metal oxide
removed from the sample by purging with an Ar flow. Both steps
composition as 68.6% CaO, 0.8% silica, 0.2% Al2O3 and 0.5% MgO,
were sequentially repeated until the adsorption of CO2 reached
with the balance as CO2. Zinc nitrate hexahydrate (Zn(NO3)26H2O)
saturation. The accumulative amount of CO2 disappearing in each
and Al2O3 were obtained as analytical reagent (AR) grade from Ajax
step, corresponding to the amount of basic sites, was detected with
Finechem. Nitric acid (HNO3; AR grade, 70 wt%) was supplied by J.
a thermal conductivity detector (TCD). The N2 physisorption
T. Baker Macron Fine Chemicals. Refined bleached deodorized
measurements were performed on a Micromeritics ASAP 2020
(RBD) palm oil (edible grade) was obtained from the Chumporn
surface area and porosity analyzer to determine the textural
Palm Oil Industry P.L.C. (Chumporn Province, Thailand). Its fatty
properties of the catalysts. The samples (80 mg) were pretreated
acid composition and basic properties are listed in the
by degassing at 300 C for 2 h. The specific surface area (SBET)
Supplementary Information (SI) Table S1. Other reagents used
was calculated according to the Brunauer-Emmett-Teller (BET)
were commercial grade (99.5%) methanol, methyl heptadecanoate
equation from the adsorption branch in the relative pressure of
(99.5%, Fluka), N-methyl-N-(trimethylsilyl)tri-fluoroacetamide
0.020.2. The total pore volume (Vp) was estimated from the
(MSTFA) (99%, Fluka), tricaprin (99%, Sigma Aldrich), pyridine
amount of N2 adsorbed at a relative pressure of about 0.990.
(99.8%, Carlo Erba) and n-heptane (99.8%, Riedel-deHan).
2.4. Transesterification procedure
2.2. Preparation of the ZSA catalyst by dissolution-precipitation
The transesterification of RBD palm oil with methanol was per-
In a typical procedure, the waste mixed seashells were ground formed in a 100-mL round bottom flask equipped with a water-
and sieved to reduce their size to less than 10 lm. After pretreating cooled condenser and a magnetic stirrer. Typically, the calcined
the seashells at 600 C for 2 h, they were dispersed in an aqueous catalyst (1 g) was vigorously stirred in methanol (11.4 g) for
solution of Zn(NO3)26H2O previously adjusted to pH 1 with 1 M 30 min, followed by the addition of the RBD palm oil (10.0 g).
HNO3. The white slurry formed was vigorously stirred at ambient The temperature of reaction mixture was maintained at 60 C
temperature for 120 min, and then Al2O3 powder was added to using a water bath. The amount of the catalyst was varied between
W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543 537

5 and 15 wt% (based on the weight of palm oil) with a constant


methanol:oil molar ratio of 30:1, whilst the methanol:oil molar
ratio was varied from 20:150:1 at a constant catalyst loading of
10 wt%. After the course of reaction (13 h), the solid phase was (e)
recovered by centrifugation at 5000 rpm. The excess methanol
was removed from the products using a rotary evaporator
(HEIDOLPH LABOROTA 4003) operating at 60 C, 0.10 atm. The
remaining liquid was washed with deionized water, resulting in (d)
two-phase product. The top layer of FAME and glyceride deriva-

Intensity/ a.u.
tives was taken, dried with anhydrous Na2SO4 and diluted with
n-heptane. The FAME composition was analyzed according to the
standard method EN 14103:2003 using a Shimadzu 14B gas chro-
(c)
matograph (GC) equipped with a flame ionization detector (FID)
and a 30-m DB-Wax capillary column. Methyl heptadecanoate
was used as the reference standard. The FAME yield was calculated
according to our previous work [24]. To quantify the mono-,
(b)
di- and tri-glycerides, MSTFA was added into the liquid sample
to convert the mono- and di-glycerides to more volatile deriva-
tives. The analysis was performed following the standard EN
14105 procedure on an Agilent 7890A GC equipped with a FID (a)
and a 15-m DB-5ht capillary column. Pyridine and tricaprin were
used as the solvent and reference standard, respectively. The GC 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
conditions for the FAME composition analysis and the glycerides 2 Theta/ degree
quantification are shown in Tables S2 and S3 (SI), respectively.
Fig. 1. XRD patterns of (a) the as-synthesized ZSA and the ZSA-x catalysts
calcined at (b) 300, (c) 500, (d) 700 and (e) 900 C. (Symbols: j = CaCO3,
H CaO, 4 = Ca(OH)2, s = Ca12Al14O33, d = ZnO, . = Zn5(OH)8(NO3)22H2O,
2.5. Reusability of the ZSA-500 catalyst h = ZnAl2O4 and = CaZn2(OH)62H2O).

To ascertain the reusability of the ZSA-500 catalyst, the spent


catalyst was recovered from the reaction mixture by centrifugation Similarly, the residual CaCO3 (2h = 29.5) in the ZSA-900 was
and reused in fresh transesterification reactions of RBD palm oil related to the adsorption of atmospheric CO2.
and methanol for five successive runs. In addition, the effect of The as-synthesized ZSA exhibited a multi-step thermal decom-
methanol washing of the spent catalyst on the subsequent FAME position pattern (Fig. 2). The first weight loss stage in the differen-
yield obtained was evaluated. For spent catalyst washing, the tial thermogravimetric (DTG) curve (<100 C) was related to the
recovered ZSA-500 catalyst was dispersed in 50 mL of fresh metha- evaporation of residual moisture, whilst the weight losses in the
nol with stirring for 15 min, followed by centrifugation to separate range of 120220 C were related to the Zn(OH)8(NO3)22H2O and
the solid phase and then adding the washed catalyst into the fresh CaZn2(OH)62H2O decomposition, according to Eqs. (1) [26] and
reaction mixture without prior drying. (2) [27], respectively.

Zn5 OH8 NO3 2  2H2 O ! 5ZnO 2HNO3 5H2 O 1


3. Results and discussion
CaZn2 OH6  2H2 O ! CaOH2 2ZnO 4H2 O 2
The weight loss found at 256 C corresponded to the dehydra-
3.1. Catalyst characterization tion of Al(OH)3 to Al2O3. Due to the complexity of the weight loss
between 300 and 600 C, the Origin 8.5 program was applied to
The XRD patterns of ZSA before and after calcination at deconvolute the DTG peaks in this temperature range (inset;
300900 C are shown in Fig. 1. The major crystalline phases of the Fig. 2). The Ca(OH)2 derived from CaZn2(OH)62H2O (Eq. (2)) was
as-synthesized ZSA were Ca(OH)2, CaCO3 and CaZn2(OH)62H2O, decomposed at 375 C, whereas the decomposition steps observed
with small diffraction peaks of Zn5(OH)8(NO3)22H2O present. The at 395 and 515 C were related to Ca(OH)2 and Ca(NO3)2, respectively.
CaCO3 was originally present in the calcined seashells, while the
other phases were generated by the dissolution-precipitation of
calcium compounds in the acidified Zn(NO3)2 solution. The 10
ZSA-300 catalyst had ZnO as new phase, whereas Zn5(OH)8(NO3)2 700 515
5
2H2O and CaZn2(OH)62H2O were absent. After calcination at 600 0
500 C, Ca(OH)2 was dehydrated to CaO, whilst the increased inten- 395
375 -5
DTG/ g min-1

sity of the CaCO3 peaks was ascribed to recarbonation of CaO 500


Weight loss/%

-10
during the catalyst preparation under the ambient atmosphere.
400
Moreover, ZnAl2O4 and Ca12Al14O33 were generated by the solid- -15
state reactions of ZnO [4] and CaO [11] with Al2O3, respectively. 300 515 -20
When the calcination temperature was increased to >700 C, the 256 -25
200 118 166 390
CaCO3 in the waste seashells was extensively decomposed to 691 -30
CaO. The enhanced peaks of CaO, ZnO, ZnAl2O4 and Ca12Al14O33 64 197
100
-35
indicated that their large crystallites were preferentially formed
0 -40
at these high temperatures. In addition, small peaks of contaminat- 0 200 400 600 800 1000
ing Ca(OH)2 (2h = 17.9, 34.1 and 47.0) were observed in the Temperature/ C
catalysts calcined at 700900 C, which was due to rehydration
of CaO with atmospheric moisture during the XRD analysis. Fig. 2. Weight loss and DTG curves of the as-synthesized ZSA.
538 W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543

The final weight loss at 691 C was due to decarbonation of CaCO3


to CaO. The phase compositions of the as-synthesized ZSA, as
determined by TGA, revealed that more than 55 wt% of the compo- 1
nents that were present as calcium compounds could be converted 2
to CaO at temperatures around 500 C (Table S4, SI).
The FTIR spectra of the ZSA-x catalysts calcined at different tem-
peratures are shown in Fig. 3, and the corresponding assignments
are summarized in Table S5 (SI). The band at 3643 cm1 was due to
m(OAH) of the Ca(OH)2 structure. The bands at 1428 and 875 cm1
were ascribed to the c(CAO) and d(CAO) of CaCO3, respectively. 3
The presence of CaZn2(OH)62H2O in ZSA-300 was evidenced by
the bands at 3575 and 3507 cm1. The band at 1348 cm1
originated from NO 3 in the structure of Ca(NO3)2. Other bands
below 940 cm1 were related to various metal-oxygen bands
(Table S5, SI). When the calcination temperature was increased
to 500 C, the bands corresponding to the OAH stretching of
CaZn2(OH)62H2O disappeared concomitantly with the absence of
O Ca
NO 3 . As deduced from a remarkable decrease in the CO3
2
band
intensities, CaCO3 was decomposed at temperatures >700 C. The 1
enhanced intensity of OAH stretching (3643 cm1) observed for
ZSA-700 was related to the rehydration of CaO to Ca(OH)2 that
was favorably formed on this catalyst (Section 3.3). The decompo- Ca
Zn Al
sition of these mixed metal compounds generated a new band at
Ca
ca. 450 cm1 that was attributed to the presence of discrete ZnO
and CaO in the lattice. Thus, the FTIR results confirmed the phase
composition of ZSA-x catalysts in agreement with the XRD and
TGA results.
The SEM image of ZSA-500, as a representative catalyst,
revealed the presence of particles with different shapes and sizes, Al Ca
O
whilst the EDS analysis of the elemental composition of the surface 2
of these particles revealed their different compositions (Fig. 4). The
small spheres with diameters in the range of 0.51.0 lm (location
No. 1 in Fig. 4), some of which were merged into or embedded in Zn
larger particles, had a Ca:Zn:Al:O atomic ratio of 11.8:1:1.8:67.4.
This was ascribed to the presence of calcium-rich compounds Ca Ca
mixed with Al2O3. The plate-like particles (location No. 2 in
Fig. 4) were Ca12Al14O33 [19], as suggested by the Ca:Zn:Al:O
atomic ratio of 5.1:1:6.6:24.3. The smaller particles (location No.
3 in Fig. 4) had a Ca:Zn:Al:O atomic ratio of 8.4:4.6:1:27.9 and so
Ca
Zn
3
(d)
O

Ca
(c) Al
Ca

(b)
%T

Fig. 4. SEM image (5000 magnification) of the ZSA-500 catalyst and the EDS
spectra from the indicated three regions of the SEM image.

(a)
were mixed oxides of Ca and Zn that would have been derived from
1045
3643

the co-precipitation of dissolved calcium species with Zn(NO3)2 in


1630

712

the solution [16]. These results indicated that the surface of the
940

570
450
822
3575

ZSA-500 catalyst was richer in calcium than the bulk composition


3507
3433

875

(Ca:Zn:Al = 2.3:1:1). It is possible that some of the calcium com-


1348
1428

pounds were dissolved from the calcined seashells and then the
dissolved calcium species (Ca2+) were reprecipitated with Zn2+
4000 3600 3200 2800 2400 2000 1600 1200 800 400
and Al2O3 to form the different Ca, Zn and Al mixed compounds.
Wavenumber/ cm-1
The physicochemical properties of the ZSA-x catalysts obtained
Fig. 3. FTIR spectra of the ZSA-x catalysts calcined at (a) 300, (b) 500, (c) 700 and after calcination at different temperatures are summarized in
(d) 900 C. Table 1. Unlike the ZSA-x catalysts calcined at temperatures
W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543 539

Table 1
Effect of the catalyst calcination temperature on the physicochemical properties and the FAME yield obtained from different ZSA-x catalysts.

Catalysta Metal phaseb Cluster size of CaOc SBETd Vpe Dpf Basicityg Initial rate FAME yieldh
(nm) (m2 g1) (cm3 g1) () (lmol g1) (gFAME g1cat. h
1
) (wt%)
ZSA-300 CC, CH, CN, ZO, AO 25.93i 5.53 0.007 47.9 28 5.74 95.2 0.85
ZSA-500 CC, CO, ZO, ZA, AO, CA 34.14 12.68 0.019 61.7 53 3.99 98.0 1.60
ZSA-700 CC, CO, ZO, ZA, AO, CA 40.81 8.88 0.035 159.8 25 1.23 95.2 2.55
ZSA-900 CO, ZO, ZA, AO, CA 50.71 5.72 0.004 27.1 20 2.76 91.4 2.60
a
Catalysts are shown as ZSA-x, where x is the calcination temperature (C).
b
Metal phase: CC = CaCO3, CH = Ca(OH)2, CN = Ca(NO3)2, CO = CaO, ZO = ZnO, ZA = ZnAl2O4, AO = Al2O3 and CA = Ca12Al14O33.
c
Determined from the XRD patterns using Sherrers equation.
d
BET surface area.
e
Total pore volume.
f
Average pore size.
g
Determined by CO2 pulse chemisorption.
h
Reaction conditions: catalyst amount, 10 wt%; palm oil amount, 10 g; methanol:oil molar ratio, 30:1; temperature, 60 C; time, 3 h.
i
Cluster size of Ca(OH)2.

>500 C, ZSA-300 possessed mainly Brnsted basic sites of nano- but can improve the mechanical and chemical stability of the
sized Ca(OH)2. From previous reports, the other metal compounds calcium-based catalysts [22].
formed at 300 C were expected to contribute a minor basicity Overall, the as-synthesized ZSA consisted of Zn5(OH)8(NO3)2
[20,21,28]. The highest surface area and basicity was observed in 2H2O, CaZn2(OH)62H2O, Ca(OH)2 and Ca(NO2)3 supported on
the ZSA-500 catalyst, and this is accounted for by the transforma- CaCO3 and Al2O3 (Scheme 1). When the catalyst was calcined at
tion of Ca(OH)2 and Ca(NO3)2 to CaO nanocrystallites dispersed on 300 C, Zn5(OH)8(NO3)22H2O and CaZn2(OH)62H2O were decom-
Al2O3. Our previous work demonstrated that the surface area and posed to give ZnO and ZnO/Ca(OH)2, respectively. At 500 C, CaO
basicity of ZSA-500 were higher than those of pure CaO obtained was generated by the dehydration of Ca(OH)2 concomitantly with
from calcination of waste mixed seashells at 800 C [14]. It can the formation of ZnAl2O4 and Ca12Al14O33 on the interfaces
be seen that the ZSA-x catalysts possessed meso-scale porosities, between the ZnO, CaO and Al2O3. The presence of both mixed oxide
which should be defined as interparticle voids similarly to other phases indicated a strong interaction of the basic oxides with the
CaO-based catalysts [5,16,18]. Further increasing the calcination Al2O3 support. By increasing the calcination temperature to 700
temperature increased the crystallite size of CaO and decreased and 900 C, the crystallite sizes of CaO and ZnO were enlarged
the total basicity, which is due to the agglomeration of CaO clusters and the formation of the mixed aluminate phases was promoted.
at high temperatures and that the majority of large CaO crystallites According to the previous works [14], we believed that Ca(OH)2
were generated directly from the decomposition of CaCO3 in the and CaO are the main active phases in the transesterification with
original seashells. Moreover, a portion of CaO was consumed in refluxed methanol, while other metal phases should have much
the formation of Ca12Al14O33 when the calcination was performed less activity due to their low basicity and mild reaction conditions
at >500 C. This calcium aluminate phase has a low basicity [12], used in the present study.

ZHN CZ CH/ CN ZnO ZnOCH CH/ CN

Al2O3/ CaCO3 300 C Al2O3/ CaCO3


HNO3
As-synthesized ZSA H2O
500 C H2O
NO, O2

ZnO ZnOCaO CaO ZnO ZnOCaO CaO


700 C
Al2O3/ CaO/ CaCO3 Al2O3/ CaCO3

ZnAl2O3 Ca12Al14O33 CO2 ZnAl2O3 Ca12Al14O33

900 C

ZnO ZnOCaO CaO

Al2O3/ CaO/ Ca 12 Al14O33

ZnAl2O3

Scheme 1. Model for the ZSA-x catalysts formed at different calcination temperatures. (Metal compounds: ZHN = Zn5(OH)8(NO3)22H2O, CZ = CaZn2(OH)62H2O,
CH = Ca(OH)2, CN = Ca(NO2)3).
540 W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543

3.2. Transesterification of RBD palm oil with methanol over the ZSA-x the CaO phase was reduced without any significant increase in the
catalysts intensities of the CaCO3 and Ca(OH)2 phases. A remarkable change
in the crystalline structure was observed in the ZSA-700 and
The ZSA-x catalysts calcined at different temperatures were ZSA-900 catalysts after the air exposure, where both had a large
tested in the catalysis of the transesterification of RBD palm oil amount of CaO as the major basic site before exposure but either
with methanol to FAME. The results are summarized in Fig. 5, no or only a small amount in the ZSA-700 and ZSA-900 catalysts,
whilst the initial rate of reaction and the maximum yield of FAME respectively, after exposure.
attained are given in Table 1. The ZSA-300 catalyst gave the fastest The chemisorption of CO2 and H2O onto CaO yields CaCO3 and
initial reaction rate, being from 1.44- to 4.67-fold faster than the Ca(OH)2, respectively. The activation energy (Ea) determined by
ZSA-x catalysts calcined at higher temperatures, which is presum- previous studies suggested that hydration of CaO (Eq. (3);
ably related to the Brnsted basicity of Ca(OH)2 and the smaller Ea = 11.020.3 kJ mol1) [30] occurs more readily than carbonation
size of Ca(OH)2 clusters. However, the slowest initial rate of reac- of CaO (Eq. (4); Ea = 2433 kJ mol1) [31]. Lanas and Alvarez
tion was obtained with the ZSA-700 catalyst and not the ZSA-900 observed the transformation of CaO to CaCO3 via generation of
one. Kouzu and co-workers reported that Ca(OH)2 was active as a Ca(OH)2 in ambient atmosphere since the carbonation at low tem-
transesterification catalyst, giving a FAME yield of 93% at 3.5 h perature requires water for CO2 dissolution on the solid surface
[10]. However, the initial rate did not proportionally correlate to [32]. The crystallinity and stability of resulting Ca(OH)2 and CaCO3
the total basicity of the catalysts, suggesting that some of the basic phases were determined by water amount and CO2 pressure.
sites were inaccessible or hindered for the bulky triglyceride mole- Besides, degree of carbonation strongly depended on crystallite
cules due to the presence of various types of metal compounds size of Ca(OH)2 [33]. Since CaCO3 (36.9 cm3 g1) has larger volume
mixed in the catalysts. The absence of a clear trend between the than Ca(OH)2 (33.6 cm3 g1) and CaO (16.9 cm3 g1) [34], the
basicity and activity of the catalysts is unsurprising since small formation of CaCO3 via direct carbonation of Ca(OH)2 (Eq. (5))
probe molecule (CO2) were used in the chemisorption measure- favors on large crystals [33].
ment [29]. The difference in the phase composition and the nature
CaOs H2 Og $ CaOH2s 3
and distribution of basic sites could account for the different FAME
yields obtained with time. Nevertheless, the ZSA-500 catalyst, with CaOs CO2g $ CaCO3s 4
the highest basicity, gave the highest FAME yield after 2 h, empha- CaOH2s CO2g $ CaCO3s H2 Og 5
sizing the beneficial influence of a high surface area and high
dispersion of CaO nanocrystallites. The maximum FAME yield The less perturbed structures of the ZSA-300 and ZSA-500 cata-
attained at 3 h was ranked in the descending order: ZSA-500 > lysts can be explained by the dispersed small active phase crystal-
ZSA-700 > ZSA-300 > ZSA-900. In addition, ZSA-500 gave a higher lites to which the Al2O3 surface strongly interacted. Although the
FAME yield than the pure CaO catalyst derived from the waste CaCO3 layer formed on the catalyst surface was expected, it was
mixed seashells [14]. present as an amorphous or low crystalline carbonate phase due
to small sizes of Ca(OH)2. In the ambient air, vaterite as metastable
CaCO3 polymorph was initially formed over Ca(OH)2 phase, and
3.3. Stability of the ZSA-x catalysts in ambient air then was transformed to more stable calcite [32]. The crystallinity
of vaterite formed was relatively low, and its presence was difficult
To reveal the stability of the ZSA-x catalysts towards the to be confirmed by XRD. The major products from the air exposure
chemisorption of CO2 and H2O present in the atmosphere, the of ZSA-700 and ZSA-900 were significantly different. ZSA-700
ZSA-x catalysts were exposed to the ambient air (RH  73%) for exhibited an extensive conversion of CaO to Ca(OH)2, whereas
3 d, and structural changes were observed by XRD analysis CaCO3 was mainly generated on the ZSA-900 exposed to air. This
(Fig. 6). When ZSA-300 was exposed to air for 3 d, it exhibited a finding was in accord with the FTIR results (Fig. 3). The larger
decreased peak intensity of Ca(OH)2, while for the ZSA-500 catalyst crystallites of CaO in ZSA-900 (Table 1) should promote the direct
carbonation of Ca(OH)2 to CaCO3 [32]. Retention of some CaO
phase in ZSA-900 after the air exposure should be ascribed to the
CaCO3 occupying the Ca(OH)2 surface as impermeable layer, which
100
prevent diffusion of CO2 and H2O into the interior crystals [35].
90 Other explanation is an enhanced interaction of CaO with Al2O3 at
80 900 C, relating to the formation of Ca12Al14O33 in the ZSA-900 catalyst.
With respect to the FAME yields attained over the different ZSA-
FAME yield / wt.%

70
x catalysts before and after the air exposure a decreased FAME
60 yield was observed in all cases (Fig. 7). For ZSA-300, the slight
50 decrease in the FAME yield of 1.4% was possibly due to the struc-
tural retention of Ca(OH)2 as an active phase. However, the FAME
40
yield decreased further with increasing calcination temperatures,
30 reaching up to a 1.18-fold or 14.2% lower yield for the ZSA-900
20 catalyst. The extensive transformation of CaO to CaCO3 is the likely
major cause of the catalytic activity loss, since the ZSA-900
10
exposed to air possessed the highest CaCO3 content. The better
0 retention of the catalytic activity (higher FAME yield) after air
0 1 2 3 exposure by the ZSA-700 catalyst than the ZSA-900 one could
Reaction time /h relate to the higher Ca(OH)2/CaCO3 phase ratio in ZSA-700
after air exposure. This result is in accord with a previous study
Fig. 5. Dependence of the FAME yield on the reaction time in the transesterification
of RBD palm oil with methanol over the ZSA-x catalysts calcined at different
that revealed a relatively large catalytic activity loss by the
temperatures. Reaction conditions: catalyst amount, 10 wt%; palm oil amount, 10 g; conversion of CaO to CaCO3 compared to the conversion of CaO
methanol:oil molar ratio, 30:1; temperature, 60 C. to Ca(OH)2 [10].
W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543 541

(A) (B)

after after

Intensity/ a.u.

Intensity/ a.u.
before before

10 20 30 40 50 60 70 80 10 20 30 40 50 60 70 80
2 Theta/ degree 2 Theta/ degree

(C) (D)
after after

Intensity/ a.u.
Intensity/ a.u.

before before

10 20 30 40 50 60 70 80 10 20 30 40 50 60 70 80
2 Theta/ degree 2 Theta/ degree

Fig. 6. Effect of air exposure on the structure change of the ZSA-x catalysts calcined at (A) 300, (B) 500, (C) 700 and (D) 900 C. (Symbols: j = CaCO3, H CaO, 4 = Ca(OH)2,
s = Ca12Al14O33, d = ZnO, . = Zn5(OH)8(NO3)22H2O, h = ZnAl2O4 and = CaZn2(OH)62H2O).

100 the calcination temperature that was used, the easier the catalysts
were separated from the reaction mixture by centrifugation
90
(data not shown). Therefore, when including the ease of catalyst
80 recovery with the catalyst stability, initial reaction rate and final
FAME yield / wt.%

FAME yield, the ZSA-500 catalyst was the selected as the most
70
suitable one for the transesterification of palm oil with methanol,
60 and so was used hereafter.
50 The effect of varying the catalyst amount and the methanol:oil
molar ratio on the FAME yield obtained over the ZSA-500 catalyst
40
in the transesterification of palm oil with methanol is shown in
30 Fig. 8. At a 30:1 methanol:oil molar ratio, the reaction rate
20 increased with increasing catalyst loading up to 10 wt% (Fig. 8A),
as expected since the amount of active sites was increased.
10 However, a catalyst loading of 15% gave a comparable FAME yield
0 after 2 h of the reaction to that with 10 wt% catalyst, suggesting
300 500 700 900 that other factors become limiting. The maximum FAME yield
Calcination temperature/ C (98.0 1.60%) was achieved at 3 h with a 10 wt% catalyst. That this
was a slightly higher FAME yield than that with a 15 wt% catalyst
Fig. 7. Effect of exposure to ambient air for 3 d on the FAME yield attained over the loading might be related to catalyst agglomeration [8,16,19]. With
ZSA-x catalysts calcined at different temperatures: ( ) before and ( ) after the a 10 wt% catalyst loading, at a 20:1 methanol:oil molar ratio, the
3-d air exposure. Reaction conditions: see Fig. 5.
FAME yield increased with time and reached 94.6 2.50 wt%
within 4 h (Fig. 8B). Increasing the methanol:oil molar ratio to
3.4. Effect of the reaction conditions on the FAME production over the 30:1 actively promoted FAME production with a faster initial reac-
ZSA-500 catalyst tion rate, and reached 98.0 1.60 wt% within 3 h. However, further
increasing the ratio to 50:1 decreased both the initial reaction rate
Generally, the recovery and regeneration of pure CaO catalysts and the final FAME yield. The effect of the methanol:oil molar ratio
is difficult due to colloid formation and the transformation of the on the FAME yield at 1 h was not as pronounced as that at 2 h. The
active phase to Ca(C3H7O3)2 that can leach out [23,25]. In the pre- amount of methanol used in this study was much higher than the
sent study, although ZSA-300 had a higher initial reaction rate and theoretical ratio, and so the excess methanol might facilitate the
stability in ambient air than the other ZSA-x catalysts, the higher removal of polar products, such as monoglycerides and glycerol,
542 W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543

100 100
90 (A) 90 (B)
80 80

FAME yield/ wt.%

FAME yield/ wt.%


70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 1 2 3 4 0 1 2 3 4
Reaction time / h Reaction time / h

Fig. 8. Effect of the (A) catalyst amount and (B) methanol:oil molar ratio on the FAME yield attained over ZSA-500 in the transesterification of RBD palm oil with methanol.
Other reaction conditions: see Fig. 5.

from the catalyst surface. However, the higher excess of methanol and then decreased only slightly to 96 wt% FAME yield after five
lowered the concentration of triglycerides on the catalyst surface, uses. It is worth noting that the pure CaO catalyst obtained from
rendering a slower rate of FAME formation. In conclusion, a suit- the waste seashells calcined at 800 C exhibited more severe activ-
able transesterification condition at 60 C for the ZSA-500 catalyst ity loss: the FAME yield was dropped from 99% to 80% at the fifth
was 10 wt% catalyst with a 30:1 methanol:oil molar ratio and a repetition [14].
reaction time of 3 h. The product analysis indicated that the The XRD analysis revealed that the structures of the fresh
contents of mono-, di- and tri-glycerides were 0.02, 0.06 and ZSA-500 and the spent ZSA-500 catalysts (after five cycles with
0.01 wt%, respectively, which are lower than the glycerides content methanol washing after each cycle) were slightly different
of biodiesel standard (0.2, 0.2 and 0.2 wt%) issued by the (Fig. S1, SI). The spent catalyst exhibited a lower intensity of most
Department of Energy Business, Ministry of Energy, Thailand. peaks than that of the fresh one because of the presence of organic
substances covering the catalyst surface. Nevertheless, the small
peaks of Ca(C3H7O3)2 were observed at 2h = 8.2 and 10.1 due to
3.5. Reusability of the ZSA-500 catalyst
the reaction of CaO with glycerol [23,25]. The amount of glycerol
and glyceride derivatives remaining in the spent ZSA-500 catalyst
The spent ZSA-500 catalyst was recovered from the reaction
was less than 2.0 wt%, as quantitatively determined by TGA
mixture by centrifugation and then, with or without methanol
(Fig. S2, SI), which was much smaller than that in the spent pure
washing, reused in fresh transesterification reactions of palm oil
CaO catalyst (13.5 wt%) [14]. Thus, the active phases of the
with methanol (60 C, 3 h, 10 wt% catalyst and a 30:1 methanol:
ZSA-500 catalyst had both a good tolerance to air exposure and a
oil molar ratio) for five successive runs to ascertain its reusability
stability against glycerol adsorption and transformation to
in terms of the obtained FAME yield (Fig. 9). Without methanol
glyceroxides. The small crystallites and strong interaction of the
washing of the catalyst between uses the FAME yield decreased
CaO species dispersed on Al2O3 may be crucial factors in the
slightly with each use from 99.2% to 87.3 wt% after four and five
improvement of catalyst stability.
uses. However, when the catalyst was washed with methanol
between uses, this decreased the loss of FAME yield such that
4. Conclusions
essentially the same activity was found for the first three uses

New heterogeneous base catalysts for the transesterification of


100 RBD palm oil with methanol were successfully prepared from
waste mixed seashells via the dissolution-precipitation method.
90
The total basicity, base type and textural properties of the catalyst
80 strongly depended on the temperature at which the catalyst was
calcined. The calcination at 500 C gave the ZSA-500 catalyst with
FAME yield/wt.%

70
the highest basicity and surface area due to the enhanced disper-
60 sion of CaO, generated from thermal decomposition of Ca(OH)2,
50 CaZn2(OH)62H2O and Ca(NO3)2, on the Al2O3. As a result, the max-
imum FAME yield of 98.0 1.60 wt% was achieved over ZSA-500
40
under the suitable transesterification conditions (catalyst load-
30 ing = 10 wt% based on the weight of oil, methanol:oil molar
20 ratio = 30:1, temperature = 60 C and time = 3 h). A strong interac-
tion of the mixed oxide nanocrystallites with the Al2O3 surface
10
improved not only the catalytic stability of ZSA-500 exposed to
0 ambient air for 3 d but also the stability against the glycerol
1 2 3 4 5 adsorption and transformation of CaO to calcium glyceroxides.
Numberof repetition
Acknowledgements
Fig. 9. Reusability of the ZSA-500 catalyst recovered without and with methanol
washing in the transesterification of RBD palm oil with methanol at 60 C for 3 h
with the catalyst loading of 10 wt% and methanol: oil molar ratio of 30:1. (Symbol: The authors are grateful to the Chumporn Palm Oil Industry
= no washing, = washing with methanol). Reaction conditions: see Fig. 5. Public Company Limited and the Thai Dolomite Company Limited
W. Jindapon et al. / Energy Conversion and Management 122 (2016) 535543 543

for donating the RBD palm oil and the waste seashell samples, [14] Jindapon W, Jaiyen S, Ngamcharussrivichai C. Al2O3-supported mixed Ca and
Zn compounds prepared from waste seashells for synthesis of palm fatty acid
respectively. This research was funded by Chulalongkorn
methyl esters. Chem Eng Commun 2015;202:15919.
University through the Ratchadapiseksompoch Endowment Fund [15] Albuquerque MCG, Azevedo DCS, Cavalcante Jr CL, Santamara-Gonzlez J,
(2013), project no. CU-56-912-EN, and the PTT Public Company Mrida-Robles JM, Moreno-Tost R, et al. Transesterification of ethyl butyrate
Limited. Financial and technical supports from the Thailand with methanol using MgO/CaO catalysts. J Mol Catal A 2009;300:1924.
[16] Ngamcharussrivichai C, Totarat P, Bunyakiat K. Ca and Zn mixed oxide as a
Research Fund (TRF) under the project no. IRG5780001 (Faculty heterogeneous base catalyst for transesterification of palm kernel oil. Appl
of Science, Chulalongkorn University), and the Center of Excellence Catal A 2008;341:7785.
on Petrochemical and Materials Technology (PETROMAT), [17] Kesic Z, Lukic I, Zdujic M, Jovalekic C, Veljkovic V, Skala D. Assessment of
CaTiO3, CaMnO3, CaZrO3 and Ca2Fe2O5 perovskites as heterogeneous base
Chulalongkorn University, are also acknowledged. The authors also catalysts for biodiesel synthesis. Fuel Process Technol 2016;143:1628.
wish to express their thanks to Dr. Robert Douglas John Butcher [18] Limmanee S, Naree T, Bunyakiat K, Ngamcharussrivichai C. Mixed oxides of Ca,
(Publication Counseling Unit, Faculty of Science, Chulalongkorn Mg and Zn as heterogeneous base catalysts for the synthesis of palm kernel oil
methyl esters. Chem Eng J 2013;225:61624.
University) for English language editing. [19] Wang BY, Li SF, Tian SJ, Feng RH, Meng YL. A new solid base catalyst for the
transesterification of rapeseed oil to biodiesel with methanol. Fuel 2013;104:
698703.
Appendix A. Supplementary material
[20] Zieba A, Pacua A, Serwicka EM, Drelinkiewicz A. Transesterification of
triglycerides with methanol over thermally treated Zn5(OH)8(NO3)2.2H2O
Supplementary data associated with this article can be found, in salt. Fuel 2010;89:196172.
the online version, at http://dx.doi.org/10.1016/j.enconman.2016. [21] Rubio-Caballero JM, Santamara-Gonzlez J, Mrida-Robles J, Moreno-Tost R,
Jimnez-Lpez A, Maireles-Torres P. Calcium zincate as precursor of active
06.012. catalysts for biodiesel production under mild conditions. Appl Catal B
2009;91:33946.
References [22] Meng YL, Tian SJ, Li SF, Wang BY, Zhang MH. Transesterification of rapeseed oil
for biodiesel production in trickle-bed reactors packed with heterogeneous Ca/
Al composite oxide-based alkaline catalyst. Bioresour Technol
[1] IEA (Internal Energy Agency). Technology Roadmap: Biofuels for Transport.
2013;136:7304.
Paris; 2011.
[23] Kouzu M, Hidaka J, Wakabayashi K, Tsunomori M. Solid base catalysis of
[2] Dennis YC, Xuan W, Leung MKH. A review on biodiesel production using
calcium glyceroxide for a reaction to convert vegetable oil into its methyl
catalyzed transesterification. Appl Energy 2010;89:108395.
esters. Appl Catal A 2010;390:118.
[3] Ma F, Hanna MA. Biodiesel production: a review. Bioresour Technol
[24] Lertpanyapornchai B, Ngamcharussrivichai C. Mesostructured Sr and Ti mixed
1999;70:115.
oxides as heterogeneous base catalysts for transesterification of palm kernel
[4] Bournay L, Casanave D, Delfort B, Hillion G, Chodorge JA. New heterogeneous
oil with methanol. Chem Eng J 2015;264(264):78996.
process for biodiesel production: a way to improve the quality and the value of
[25] Len-Reina L, Cabeza A, Rius J, Maireles-Torres P, Alba-Rubio AC, Lpez-
the crude glycerin produced by biodiesel plants. Catal Today 2005;106:
Granados M. Structural and surface study of calcium glyceroxide, an active
190201.
phase for biodiesel production under heterogeneous catalysis. J Catal
[5] Ngamcharussrivichai C, Nunthasanti P, Tanachai S, Bunyakiat K. Biodiesel
2013;300:306.
production through transesterification over natural calciums. Fuel Process
[26] Biswick T, Jones W, Pacu A, Serwicka E, Podobinski J. The role of anhydrous
Technol 2010;91:140915.
zinc nitrate in the thermal decomposition of the zinc hydroxy nitrates Zn5
[6] Chen GY, Shan R, Yan BB, Shi JF, Li SY, Liu CY. Remarkably enhancing the
(OH)8(NO3)22H2O and ZnOHNO3H2O. J Solid State Chem 2007;180:11719.
biodiesel yield from palm oil upon abalone shell-derived CaO catalysts treated
[27] Wang S, Yang Z, Zeng L. Study of calcium zincate synthesized by solid-phase
by ethanol. Fuel Process Technol 2016;143:1107.
synthesis method without strong alkali. Mater Chem Phys 2008;112:6036.
[7] Yin X, Duan X, You Q, Dai C, Tan Z, Zhu X. Biodiesel production from soybean
[28] Suppes GJ, Bockwinkel K, Lucas S, Mason JB, Heppert JA. Calcium carbonate
oil deodorizer distillate using calcined duck eggshell as catalyst. Energy
catalyzed alcoholysis of fats and oils. J Am Oil Chem Soc 2001;78:13945.
Convers Manage 2016;112:199207.
[29] Ryczkowski J. IR spectroscopy in catalysis. Catal Today 2001;68:263381.
[8] Syazwani ON, Rashid U, Yap YHT. Low-cost solid catalyst derived from waste
[30] Wang Y, Lin S, Suzuki Y. Effect of CaO content on hydration rates of Ca-based
Cyrtopleura costata (Angel Wing Shell) for biodiesel production using
sorbents at high temperature. Fuel Process Technol 2008;89:2206.
microalgae oil. Energy Convers Manage 2015;101:74956.
[31] Sun P, Grace JR, Lim J, Anthony EJ. Determination of intrinsic rate constants of
[9] Granados ML, Poves MDZ, Alonso DM, Mariscal R, Galisteo FC, Moreno-Tost R,
the CaOCO2 reaction. Chem Eng Sci 2008;63:4756.
Santamara J, Fierro JLG. Biodiesel from sunflower oil by using activated
[32] Lanas J, Alvarez JI. Dolomitic limes: evolution of the slacking process under
calcium oxide. Appl Catal B Environ 2007;73:31726.
different conditions. Thermochim Acta 2004;423:112.
[10] Kouzu M, Kasuno T, Tajika M, Yamanaka S, Sugimoto Y, Yamanaka S, et al.
[33] Blamey J, Lu DY, Fennell PS, Anthony EJ. Reactivation of CaO-based sorbents for
Calcium oxide as a solid base catalyst for transesterification of soybean oil and
CO2 capture: mechanism for the carbonation of Ca(OH)2. Ind Eng Chem Res
its application to biodiesel production. Fuel 2008;87:2798806.
2011;50:1032934.
[11] Meng YL, Wang BY, Li SF, Tian SJ, Zhang MH. Effect of calcination temperature
[34] Perry RH, Green DW. Perrys chemical engineerss handbook. 7th ed. New
on the activity of solid Ca/Al composite oxide-based alkaline catalyst for
York: McGraw-Hill; 1997.
biodiesel production. Bioresour Technol 2013;128:3059.
[35] Montes-Hernandez G, Pommerol A, Renard F, Beck P, Quirico E, Brissaud O. In
[12] Benjapornkulaphong S, Ngamcharussrivichai C, Bunyakiat K. Al2O3-supported
situ kinetic measurements of gas-solid carbonation of Ca(OH)2 by using an
alkali and alkali earth metal oxides for transesterification of palm kernel oil
infrared microscope coupled to a reaction cell. Chem Eng J 2010;161:2506.
and coconut oil. Chem Eng J 2009;145:46874.
[13] Liu Y, Zhang P, Fan M, Jiang P. Biodiesel production from soybean oil catalyzed
by magnetic nanoparticle MgFe2O4@CaO. Fuel 2016;164:31421.

You might also like