You are on page 1of 12

Energy Conversion and Management xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Evaluation of palm oil mill fly ash supported calcium oxide


as a heterogeneous base catalyst in biodiesel synthesis
from crude palm oil
Wilson Wei Sheng Ho a, Hoon Kiat Ng a, Suyin Gan b,⇑, Sang Huey Tan a
a
Department of Mechanical, Materials and Manufacturing Engineering, The University of Nottingham Malaysia Campus, Jalan Broga, 43500 Semenyih, Selangor Darul Ehsan, Malaysia
b
Department of Chemical and Environmental Engineering, The University of Nottingham Malaysia Campus, Jalan Broga, 43500 Semenyih, Selangor Darul Ehsan, Malaysia

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

Article history: A palm oil mill fly ash supported calcium oxide (CaO) catalyst was developed to be used as a heteroge-
Available online xxxx neous base catalyst in biodiesel synthesis from crude palm oil (CPO). The catalyst preparation procedure
was optimised in terms of final calcination temperature and duration. The optimum catalyst preparation
Keywords: conditions were determined as final calcination at 850 °C for 2 h with 45 wt.% loading of calcined calcium
Crude palm oil carbonate (CaCO3). A maximum biodiesel yield of 75.73% was achieved for this catalyst under fixed
Heterogeneous catalyst transesterification conditions. Characterisation tests showed that the catalyst had higher surface area
Transesterification
and basic sites which favoured transesterification. The effects of catalyst loading, methanol to oil molar
Biodiesel
Catalyst reusability
ratio, reaction temperature and reaction time on biodiesel yield and fatty acid methyl ester (FAME) con-
version were also investigated. It was determined that transesterification conditions of 6 wt.% catalyst
loading, 12:1 methanol to oil molar ratio, 45 °C reaction temperature, 3 h reaction time and 700 rpm stir-
ring speed resulted in biodiesel yield and FAME conversion of 79.76% and 97.09%, respectively. Experi-
mental kinetic data obtained from the heterogeneous transesterification reactions fitted the pseudo-
first order kinetic model. The activation energy (Ea) of the reaction was calculated to be 42.56 kJ mol1.
Key physicochemical properties of the produced biodiesel were measured and found to be within the
limits set by EN 14214. The developed catalyst could feasibly be used up to three consecutive cycles after
regeneration using methanol washing followed by recalcination at 850 °C for 2 h.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction using heterogeneous catalysed transesterification include less


wastewater generation and the ability to recycle or reuse the
Heterogeneous catalysed transesterification has recently been catalyst [1]. These advantages could potentially lead to lower pro-
proposed as an alternative process to homogeneous catalysed duction costs. Nonetheless, there are two key limitations in the use
transesterification for biodiesel production. The advantages of of heterogeneous catalysts. Firstly, heterogeneous catalysts have
lower reaction rates as compared to homogeneous catalysts. Sec-
Abbreviations: Al, aluminium; Al2O3, aluminium oxide; ASTM, American Society ondly, the catalyst surface active sites are rapidly poisoned upon
for Testing and Materials; BET, Brunauer–Emmett–Teller; Ca, calcium; CaCO3, exposure to ambient conditions because of chemisorption of car-
calcium carbonate; CaO, calcium oxide; Ca(OH)2, calcium hydroxide; Ca2SiO4, bon dioxide (CO2) and water (H2O) onto the surface sites [2]. Thus,
dicalcium disilicate; CN, cetane number; CO2, carbon dioxide; CPO, crude palm oil; further development of heterogeneous catalysts should focus on
DG, diglyceride; DOBI, bleachability index; Ea, activation energy; EDS, energy
dispersive X-ray spectroscopy; FAME, fatty acid methyl ester; FFA, free fatty acid;
enhancing not only catalytic performance, but also its reusability.
FID, flame ionization detector; GC, gas chromatography; HHV, high heating value; Calcium oxide (CaO) has recently been reported as an active het-
H2O, water; ICP-AES, inductively coupled plasma-atomic emission spectroscopy; erogeneous base transesterification catalyst due to its high basicity,
JCO, Jatropha curcas oil; K, potassium; k, reaction rate constant; ko, pre-exponential low solubility and easy handling [3–6]. The use of CaO can be cate-
factor; Li/CaO, lithium doped calcium oxide; MG, monoglyceride; Mg, magnesium;
gorised into five main groups. The groups are neat CaO, doped CaO,
N2, nitrogen; O, oxygen; P, phosphorus; R, gas constant; R2, correlation coefficient; r,
reaction rate; SEM, scanning electron microscopy; Si, silicon; SiO2, silicon oxide; T, mixed CaO, waste CaO and supported CaO [4]. With regard to neat
temperature; t, time; TG, triglycerides; TGA, thermogravimetric analysis; XME, CaO applications, Kouzu et al. [7] has examined the reactivity of
conversion of methyl ester; XRD, X-ray diffraction; ZnO, zinc oxide. CaO, calcium hydroxide (Ca(OH)2) and calcium carbonate (CaCO3)
⇑ Corresponding author. Tel.: +60 3 89248162; fax: +60 3 89248017.
in converting soybean oil into biodiesel. Fatty acid methyl ester
E-mail address: suyin.gan@nottingham.edu.my (S. Gan).

http://dx.doi.org/10.1016/j.enconman.2014.03.061
0196-8904/Ó 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
2 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

(FAME) conversions of 93%, 12% and 0% were reported for CaO, index (DOBI) value of 2.6% and a moisture content of 0.2%. Palm oil
Ca(OH)2 and CaCO3, respectively. However, it was reported that mill fly ash was collected from Seri Ulu Langat Palm Oil Mill Sdn.
the catalytic activity of CaCO3 could be improved by employing Bhd., Malaysia to be used as catalyst support. Methanol (>99% pur-
thermal activation treatment such as calcination in order to remove ity) and CaCO3 (99.95% purity) were purchased from Merck and
the surface carbonates [8]. To enhance the performance of CaO, Sigma Aldrich, respectively. Meanwhile, an internal standard
some researchers have tried doping an active ingredient onto CaO methyl heptadecanoate was purchased from Sigma Aldrich.
[9,10]. For instance, lithium doped CaO (Li/CaO) was examined by
Alonso et al. [11] in the transesterification of sunflower oil. The 2.2. Catalyst synthesis
authors indicated that activation of the catalyst at high temperature
was essential. However, when the activation temperature exceeded The wet impregnation method was utilised to prepare the cat-
500 °C, lixiviation occurred which affected the catalytic perfor- alyst at a fixed composition of 45 wt.% calcined CaCO3 loaded onto
mance. CaO has also been mixed with other oxides for transesteri- palm oil mill fly ash [17]. In brief, the fly ash was first sieved and
fication. It has been reported that CaO mixed with zinc oxide (ZnO) dried at 105 °C for 24 h to eradicate all moisture. Prior to use,
used in the transesterification of palm kernel oil to biodiesel re- CaCO3 was calcined at 800 °C for 90 min to remove CO2 and gener-
sulted in a high FAME conversion of 94% [12]. In another mixed ate CaO. Subsequently, water was added to obtain the active salt
CaO attempt, Zhu et al. [13] treated CaO with ammonium carbonate precursor Ca(OH)2. The purpose of dissolving CaO into aqueous
solution and calcined it at 900 °C to form a solid super base, which Ca(OH)2 was to create a suitable medium for both the CaO and
showed high catalytic activity in Jatropha curcas oil (JCO) transeste- fly ash to combine well during mixing. Typically, to obtain
rification. Under the optimum conditions, the conversion of JCO 45 wt.% calcined CaCO3 loaded catalyst, 13.5 g of calcined CaCO3
reached 93%. In order to develop more cost effective CaO based cat- was added to 200 mL of pure deionised H2O in a 3-neck flask under
alysts, waste CaO and supported CaO are preferred. Waste mud magnetic stirring at 700 rpm to prepare an aqueous solution of
crab, cockle and chicken eggs shells were reportedly utilised as Ca(OH)2. The solution was heated in a closed system at 70 °C to
CaO based catalysts in palm olein transesterification [14–16]. minimise water loss. 16.5 g of fly ash was then slowly added to
Previously, CaO supported on different palm oil mill waste boiler the solution and mixed vigorously for another 4 h until a homoge-
ashes were developed and characterised by the authors for the nous mixture was obtained. The sample was then aged for 18 h to
transesterification of CPO [17]. It was found that the use of CaO sup- ensure that the Ca(OH)2 precipitated on the fly ash carrier. The re-
ported on waste fly ash resulted in the best FAME conversion of moval of adhering H2O in the mixture was accomplished using a
94.5% under fixed transesterification conditions [17]. The present hot air oven at 105 °C for 24 h. A hot filtration test following the
study focuses on further developing and evaluating the waste fly methodology given by Lempers and Sheldon [25] was carried out
ash supported CaO as a catalyst to convert CPO to biodiesel. Four as- to confirm that the developed catalyst was heterogeneous. Briefly,
pects are covered in this study, all of which have not been reported the catalyst was filtered from the transesterification reaction mix-
in the literature before. Firstly, the wet impregnation method used ture halfway through the reaction at 2 h and 80 °C (6 wt.% catalyst
in catalyst development is optimised in terms of calcination tem- loading and 12:1 methanol to oil molar ratio) and the reaction was
perature and duration after metal loading. The motivation for this allowed to continue in the absence of the catalyst. The contribution
optimisation is to increase the catalytic performance and poten- of homogeneity was ruled out since the reaction came to a halt.
tially reduce the energy consumption in catalyst preparation. Sec- To develop a catalyst with increased catalytic activity and
ondly, parametric transesterification experimental tests are potentially reduce the energy consumption in catalyst preparation,
carried out to elucidate the effects of catalyst loading, methanol the wet impregnation method used was optimised in terms of cal-
to oil molar ratio, reaction time and temperature on FAME conver- cination temperature and duration after metal loading. The dried
sion and biodiesel yield. For this study, both FAME conversion and catalyst was calcined at temperatures of 850, 900, 950 and
biodiesel yield are measured to evaluate the reaction process. In the 1000 °C in a muffle furnace for two different durations of 2 and 5 h.
majority of the reported papers on transesterification, the effective-
ness of the reaction is solely judged by the FAME conversion 2.3. Catalyst characterisation
[2,8,14,15,18–20]. Although biodiesel yield is important as an indi-
cator of the process viability, it is seldom measured. A minority of The basic strength of the catalysts was determined by using the
studies have measured the biodiesel yield [13,21] or both FAME following Hammett indicators: phenolphthalein (H_ = 9.3), 2,4-
conversion and yield [22,23] in order to evaluate the transesterifi- dinitroaniline (H _ = 15) and 4-nitroaniline (H_ = 18.4). Typically,
cation reaction. The focus of the parametric tests is to determine 25 mg of catalyst was mixed with 1 mL of Hammett indicator di-
the conditions which would result in the highest possible biodiesel luted in 20 mL of methanol and equilibrated for 2 h. A colour
yield without compromising the FAME conversion which must ex- change indicated that the catalyst is stronger than the indicator
ceed 96.5% as specified in the European Standard EN 14214 [24]. and vice versa [14]. The synthesised catalysts were examined by
Under optimised conditions, the key biodiesel properties are also thermogravimetry analysis (TGA) using a Mettler Toledo TGA/
determined. Thirdly, the transesterification kinetic using the waste DSC 1 system operating from ambient temperature to 900 °C at a
fly ash supported CaO catalyst is studied in order to determine the scanning rate of 10 °C min1 under a nitrogen (N2) flow rate of
reaction rate constant (k) and activation energy (Ea). Lastly, the 15 mL min1. The surface morphology of the developed catalysts
reusability of this heterogeneous base catalyst is examined in light was analysed using scanning electron microscopy (SEM) equipped
of the need for development of enhanced reusability of heteroge- with energy dispersive X-ray spectroscopy (EDS). The catalysts
neous catalysts for biodiesel synthesis. were photographed using SEM in a FEI Quanta 400F FESEM system
operating at 15 kV to estimate the particle sizes and shapes. The
2. Materials and methods EDS elemental composition (particularly the calcium (Ca) concen-
tration) was semi-quantitatively expressed in terms of weight per-
2.1. Materials and chemicals centages using Oxford-instruments INCA 400. EDS was used due to
the unavailability of more accurate quantification techniques such
CPO was provided by Havys Oil Mill Sdn. Bhd., Malaysia. The as inductively coupled plasma-atomic emission spectroscopy (ICP-
accompanying sample analysis report stated that the CPO had a AES). In addition, X-ray diffraction (XRD) was utilised as a quanti-
free fatty acid (FFA) content of 3.7%, a deterioration of bleachability tative determination of the CaO concentration within the catalysts

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx 3

[26]. The crystalline structure of the catalysts was also character- Weight of biodiesel
Yield ð%Þ ¼ ð1Þ
ised via XRD in a PANanalytical X’pert Pro system using Cu Ka radi- Weight of oil
ation (40 mA and 45 kV) equipped with a Xe proportional detector.
The analysis was performed at a scanning rate of 1° min1 (step
2.5. Experimental kinetic study
size of 0.02°) in the 2h range from 10° to 80°. The specific surface
areas of the fresh and used catalysts were measured using the Bru-
Transesterification experiments were carried out at different
nauer–Emmett–Teller (BET) surface area analysis. The catalysts
reaction temperatures (45, 50, 55, 60, 65 °C) under fixed catalyst
were tested for their surface areas and porosities using a Microm-
loading of 6 wt.% and methanol to oil molar ratio of 12:1. Each
eritics’ ASAP 2020 porosimetry analyser. This was accomplished
experiment was allowed to run until 6 h to ensure complete con-
using N2 adsorption–desorption isotherms at liquid N2 tempera-
version of CPO into FAME. At every 1-h interval, a 2 ml sample
ture and relative pressures (P/Po) ranging from 0.07 to 0.3 whereby
was withdrawn from the reaction mixture for gas chromatography
a linear relationship was maintained. Prior to each measurement,
(GC) analysis. The withdrawn sample was immediately quenched
the sample was degassed at 350 °C for 3 h.
with 4 ml 0.1 M aqueous acetic acid to neutralise the reaction. It
was then centrifuged at 1600 rpm for 15 min to form a two-phase
2.4. Catalytic activity evaluation of palm oil mill fly ash supported CaO
mixture comprising an ester phase and an aqueous phase. An
catalyst
emulsion was formed in between these two phases and its thick-
ness varied proportionally to the amount of diglyceride (DG) and
The palm oil mill fly ash supported CaO catalysts prepared un-
monoglyceride (MG) formed in the reaction mixture [27]. The
der different final calcination conditions were evaluated in terms
aqueous layer was discarded while the upper ester layer was dried
of their catalytic activity in the transesterification of CPO. Fixed
in an oven at 105 °C for 24 h before GC analysis.
transesterification conditions of methanol to oil molar ratio of
12:1, temperature of 60 °C and stirring speed of 700 rpm for 3 h
were chosen [17]. The catalyst loading was varied from 2 to 2.6. Gas chromatography (GC) analysis
8 wt.% and the biodiesel yield was measured.
A single factor parametric study was then carried out to inves- The FAME content in the biodiesel samples was quantitatively
tigate the effects of catalyst loading, methanol to oil molar ratio, analysed by GC according to European Standard EN 14103 [28].
reaction time and reaction temperature on biodiesel yield and The analysis was carried out using a Perkin Elmer Clarus 500 GC
FAME conversion using the catalyst prepared under optimum con- equipped with a flame ionisation detector (FID). The capillary col-
ditions. Table 1 shows the nominal parametric settings for all the umn used was a highly polar Agilent J&W HP-88 cyanopropyl col-
transesterification tests along with the variation in each of the umn with dimensions of 0.25 mm I.D.  60 m length  0.2 lm film
parameters for the single factor parametric study. All experiments thickness. Helium was used as the carrier gas. The pressure of the
were repeated thrice. gas flow was set at 32 psi. The oven temperature was isothermally
The transesterification experiments were carried out in a held at 185 °C for 15 min. Meanwhile, the injector and detector
250 mL three-neck round bottom flask equipped with a reflux con- temperatures were set at 240 °C and 250 °C, respectively.
denser. A thermometer was fitted onto one side-neck while the 50 mg of sample was placed in a 2 mL vial and 1 mL of internal
other side-neck was used for sampling. The reactor was placed in standard methyl heptadecanoate solution (10 mg/mL) was added
a heating mantle with controlled heating and stirring. For all tests, using a pipette. The methyl heptadecanoate solution was prepared
100 mL (87 g) of CPO was used. The CPO was first heated to 5 °C by weighing 500 mg of methyl heptadecanoate in 50 mL of hexane.
above the desired reaction temperature. The desired amount of 1 lL of sample was injected into the GC with a split injection ratio
fly ash supported CaO catalyst was added to the methanol and of 100:1. This quantitative analysis provides verification that the
was premixed at 40 °C and 500 rpm for 15 min. The catalyst-meth- ester content in biodiesel is greater than 96.5% m/m in accordance
anol mixture was then introduced into the reactor. Stirring was ini- with the specifications reported in EN 14214 [24]. As detailed in
tiated and the reaction was carried out at the desired temperature the European standard EN 14103, the ester content (FAME conver-
and time. sion) expressed as a mass fraction in percentage was calculated
After the reaction, the solid catalyst was filtered out from the using Eq. (2) [28].
mixture with the aid of a vacuum pump. The filtered sample was   
RA  AIS C IS V IS
separated by centrifugation at 4000 rpm for 15 min. Two phases Ester content ð%Þ ¼ 100 ð2Þ
AIS m
would be formed, namely an ester-rich phase (top layer) and a P
glycerol-rich phase (bottom layer). The glycerol was removed where A = sum of area of all peaks ranging from C14:0 and C24:1;
using a separating funnel while the excess methanol in the ester- AIS = internal standard (methyl heptadecanoate) peak area;
rich phase was removed via a rotary evaporator. The sample from CIS = concentration of the internal standard solution, in mg/mL;
the ester-rich phase underwent diluted acetic acid washing (90 °C) VIS = volume of the internal standard solution used, in mL; m = mass
to remove all dissolved residues and hot air oven drying (65 °C) to of the sample, in mg.
remove all H2O content. The mass of the final product, i.e. biodiesel
was measured for the determination of yield according to Eq. (1). 2.7. Biodiesel physicochemical properties determination

Key physicochemical properties of the biodiesel produced un-


der fixed transesterification conditions of 6 wt.% catalyst loading,
Table 1
The nominal and tested parametric settings for the transesterification study.
12:1 methanol to oil molar ratio, 45 °C temperature, 3 h time and
700 rpm stirring speed were determined according to American
Reaction parameters Tested settings Society for Testing and Materials (ASTM) standards. The measured
Catalyst loading (wt.%) 2, 3, 4, 5, (6)a, 7, 8 properties included: (i) kinematic viscosity at 40 °C determined
Methanol to oil molar ratio (mol/mol) 6:1, 9:1, (12:1)a, 15:1, 18:1 using glass capillary viscometers according to ASTM D 445; (ii)
Reaction time (h) (3)a, 4, 5, 6, 7
density measured using the hydrometer method according to
Reaction temperature (°C) 45, 50, 55, (60)a, 65
ASTM D 1298; (iii) flash point measured using the rapid equilib-
a
The settings in parentheses are the nominal settings. rium closed cup method according to ASTM D 93; (iv) acid value

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
4 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

determined by volumetric titration according to ASTM D 664 and; 2.9. Catalyst reusability
(v) higher heating value (HHV) estimated using a bomb calorime-
ter accordingly to ASTM D 4868. Due to the unavailability of cetane After the first optimised transesterification cycle, the recovered
number (CN) testing facility on-campus, Eq. (3) which was devel- catalysts were tested with 3 different catalyst regeneration meth-
oped using Artificial Neural Networks (ANN) by Cheenkachorn ods to determine which method provided the best recovery of CaO
[29] was used to calculate CN: and Ca concentrations. The regeneration methods were as follows:
(i) methanol washing; (ii) recalcination without methanol wash-
CN ¼ 33:6 þ 0:539 ðC18 : 0Þ þ 0:303 ðC18 ing; and (iii) methanol washing followed by recalcination. In the
: 1Þ þ 0:0878 ðC18 : 2Þ þ 0:233 ðC22 : 1Þ ð3Þ first method, after completion of reaction, the spent solid catalyst
was separated from the mixture by filtration and washed with
where CN indicates the cetane number while (C18:0), (C18:1), methanol to remove polar compounds on its surface such as glyc-
(C18:2), (C22:1) are the mass concentrations of stearic, oleic, lino- erol. In the second method, the separated catalyst was directly
leic and erucic acids, respectively. recalcined at 850 °C for 2 h without any methanol washing. Lastly,
in the third method, the separated catalyst was washed with meth-
2.8. Kinetic modelling anol first and then dried in the oven (65 °C for 3 h) before recalcin-
ing at 850 °C for 2 h. A total of 5 transesterification cycles were
Considering that excess methanol was used in the reaction to carried out using the optimum regeneration method to study the
shift the reaction equilibrium towards the product side, the reverse catalyst reusability. The determination of the CaO and Ca concen-
reaction can be ignored [30]. To simplify the kinetic modelling, the trations of the catalyst after each regeneration method and each
reaction was assumed to be pseudo-first order reaction and hence transesterification cycle was carried out using XRD and EDS,
the following rate equations as used by Birla et al. [31] were respectively.
applicable:

d½TG 3. Results and discussion


r¼ ¼ k½TG ð4Þ
dt
3.1. Characterisation of the catalysts
where r is the transesterification reaction rate (mol mL1 min1), k
is the equilibrium rate constant (min1) and [TG] is concentration of
3.1.1. Basic strength analysis
triglycerides (mol mL1) and t is the time (min).
The catalysts changed the colour of phenolphthalein from col-
Integration of Eq. (3) from t = 0 and [TG] = [TG0] till t = t and
ourless to pink, but failed to change the colour of 2,4-dinitroaniline
[TG] = [TG] gives:
and 4-nitroaniline. This indicated that the catalysts’ basic strength
ln½TG0   ln½TGt  ¼ kt ð5Þ was within 9.3 < H_ < 15. As such, the developed catalysts were
considered as a relatively strong base catalyst for transesterifica-
From mass balance, tion although the strength was slightly lesser than the original
½TG CaO (typically 15 < H_ < 18.4 [14]).
X ME ¼ 1  ð6Þ
½TG0 

or, 3.1.2. Thermogravimetry Analysis (TGA)


Fig. 1 depicts the TGA profiles of the catalysts produced under
½TG ¼ ½TG0 ð1  X ME Þ ð7Þ different final calcination temperatures (850–1000 °C) and dura-
tions (2 and 5 h). For all temperatures and durations, only a max-
where XME is the conversion of methyl ester. Eq. (7) can also be ex- imum of 1.3% weight loss was detected. This demonstrates that
pressed in terms of the following rate equation: the final calcination process has a significant effect on the thermal
stability of the catalyst. For the 2-h calcined catalysts, the thermal
dX ME
¼ kð1  X ME Þ ð8Þ stability increased (i.e. weight loss was minimised) as the calcina-
dt
tion temperature was increased from 850 to 950 °C. An increase in
Integration and rearrangement of Eq. (8) results in: the calcination temperature enhances decomposition rate and
activity such as CaCO3 decomposition into active CaO which occurs
 lnð1  X ME Þ ¼ kt ð9Þ
between 825 and 900 °C. As reported previously, CaO displayed a
By fitting the experimental data to Eq. (9), the rate constants k high thermal stability at any temperature due to its unique chem-
at different temperatures can be obtained from the linear slope ical bonds [17]. Thus, increasing the calcination temperature above
through the origin. The activation energy Ea and pre-exponential 850 °C could improve the catalyst thermal stability. However, an
factor k0 for the transesterification reaction can be calculated using increase in final calcination temperature to 1000 °C caused a slight
the following Arrhenius equation: reduction in thermal stability. Calcination temperatures that are
Ea
too high could possibly damage the residual hydroxyl groups
k ¼ k0 eRT ð10Þ bonded to the oxide lattice and cause defects in the catalyst struc-
1
ture [32]. For the 5-h calcined catalysts, an inconsistent observa-
where Ea is the activation energy (kJ mol ), R is the gas constant tion is seen. The thermal stability did not exhibit an increasing
(8.314 J mol1 K1), T is the temperature (K), and k0 is the pre-expo- trend with temperature. The observed trend followed the order
nential factor (min1). of 950 > 1000 > 850 > 900 °C. This inconsistent trend could be due
By taking the natural logarithm of Eq. (10), Eq. (11) is obtained: to calcination duration being more dominant than temperature.
Ea 1 When calcination duration is prolonged, the thermal stability
ln k ¼   þ ln k0 ð11Þ reaches the maximum thermal stability regardless of the tempera-
R T
ture used. Although an inconsistent trend is observed, it must be
where Ea and k0 can be determined from the slope and intercept, borne in mind that the weight loss differences between each of
respectively. the cases were negligible (<0.2%).

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx 5

(a) (b)

Fig. 1. TGA profiles of calcined catalysts for (a) 2 h and (b) 5 h.

Comparing between the two different calcination durations, for result, the surface areas of the catalyst increased. For the 5-h cal-
temperatures 850 and 950 °C, the TGA profiles obtained were al- cined catalysts as presented in Fig. 2(e)–(h), the morphological
most identical. Only at 1000 °C, the increase in calcination duration sizes remained constant with the increase in calcination tempera-
from 2 to 5 h promoted thermal stability as the weight loss was re- ture. Again, as with the TGA profiles, the effect of calcination dura-
duced by 1%. Generally, the increase in calcination duration did not tion was more prominent than the effect of calcination
significantly improve the catalyst stability. However, due to the temperature. Calcination at high temperature for prolonged dura-
longer calcination process, the energy consumption increased. tion could possibly damage the agglomerated structure of the cat-
The improvement in catalyst stability observed was minimal, with alyst. This can be seen from Fig. 2(h) which shows that the catalyst
a maximum weight loss of 0.4%. Therefore, from the TGA profiles, it calcined at 1000 °C for 5 h has bigger unevenly formed and distrib-
could be concluded that calcination at 850 °C for 2 h was sufficient uted crystal chunks structure instead of a finer and more uniform
to produce a high thermal stability catalyst. structure. This observation agrees well with Rashidi et al. [33] who
reported that too high a calcination temperature can damage the
3.1.3. Scanning electron microscopy (SEM) and energy dispersive X- catalyst structure. Once more, the increase in calcination duration
ray spectroscopy (EDS) does not significantly improve the morphological characteristics of
Fig. 2 shows the morphological characteristics of the calcined the calcined catalyst, and therefore, calcination for 2 h was deemed
catalysts under different temperatures and durations as captured sufficient to produce a highly adsorbent catalyst. However, a finer
by the SEM micrographs. In general, all the catalysts exhibited sim- morphology size could be observed as the calcination temperature
ilar spongy and porous structures indicating highly adsorbent sur- increased to 950 °C. Overall, the SEM micrographs indicated that
faces. All particles shown have uniform distribution of miniature 950 °C was the most suitable calcination temperature to obtain
agglomerates with irregular shapes. These agglomerated structures the finest porous catalyst.
of the calcined catalysts indicated the formation of metal oxides Table 2 summarises the EDS elemental analysis for the devel-
upon heat treatment [21]. As shown in Fig. 2(a)–(d), for the 2-h cal- oped catalysts. Since the data presented are considered semi-quan-
cined catalysts, the morphological sizes of the particles reduced as titative, all associated discussion using the elemental weight
the calcination temperature increased from 850 to 1000 °C. As a percentages are relying on the figures to obtain valid trends when

(a) (b) (c) (d)

50 µm 50 µm 50 µm 50 µm

(e) (f) (g) (h)

50 µm 50 µm 50 µm 50 µm

Fig. 2. SEM images (2000 magnification) of catalysts calcined at (a) 850 °C, (b) 900 °C, (c) 950 °C and (d) 1000 °C for 2 h; (e) 850 °C, (f) 900 °C, (g) 950 °C and (h) 1000 °C for
5 h.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
6 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

parameters are changed. The actual magnitudes must be treated


with a certain degree of conservatism. The analysis shows the
abundance of both Ca and oxygen (O) in CaCO3 before and after
Calcined
catalyst
final calcination. The Ca concentration increased upon calcination,
1000 °C

2.685
with more than 50 wt.% Ca being detected when CaCO3 was
5h
11
42

39
68
1
1
6


calcined at 800 °C. Furthermore, the elemental analysis for the
developed catalysts indicated that Ca was adsorbed by the ashes,
as high concentrations of Ca ranging from 34 to 38 wt.% were
Calcined
catalyst
1000 °C

noticeable in the catalysts. Several other metals, namely potassium

1.443
(K), magnesium (Mg), aluminium (Al), phosphorus (P) and silicon
2h
21
39

35
60

1
4


(Si) are the major components of the ashes. For the longer calcina-
tion duration of 5 h, a general increasing trend in Ca concentration
was observed as the temperature increased. Conversely, for the
Calcined
catalyst

shorter calcination duration of 2 h, a reverse trend was seen. In


950 °C

3.399
short, the final calcination conditions of 900 and 1000 °C for 5 h re-
5h
11
44

37
63
1
1
6

sulted in the highest Ca concentrations of 39 wt.% for both cases.


Nevertheless, the results also indicated that the catalyst calcined
at 850 °C for 2 h had comparatively high Ca concentration
Calcined
catalyst
950 °C

(38 wt.%) with respect to the catalysts developed under higher


5.026
2h

calcination temperatures and longer durations.


13
43

35
71
1
1
7

3.1.4. X-ray Diffraction (XRD)


The powder XRD patterns of the developed catalysts are de-
Calcined
catalyst
900 °C

5.753

picted in Fig. 3(a)–(h). Fig. 3(a) reveals only the CaO crystalline
5h

45

39
78

peaks indicating that the catalyst developed under this condition


8

1
1
6

had high CaO content. The characteristic peaks are observed at


2h = 32.32°, 37.48°, 53.95°, 64.31° and 67.47°. Peaks due to mullite
Calcined

(mainly aluminium oxide (Al2O3) and silicon oxide (SiO2)) and


catalyst
900 °C

4.754

dicalcium disilicate (Ca2SiO4) can be seen in Fig. 3(b)–(h). These


2h

47

36
75
7

1
1
8

combined peaks are due to the strong impregnation of active


CaO within the palm oil mill fly ash. The mullite peaks representing
the major ash components are observed at 2h = 26.7°, 39.5° and
Calcined
catalyst

50.7°. Meanwhile, the Ca2SiO4 crystalline phase is observed at


850 °C

6.399

2h = 30.9° for all catalysts except for the catalyst calcined at


5h
13
46

34
73
1
1
5

850 °C for 2 h and 900 °C for 5 h. The presence of Ca2SiO4 suggested


that Ca2SiO4 hydrate formed through the reaction between SiO2
and CaO in the presence of water. The Ca2SiO4 is an impurity and
Calcined
catalyst
850 °C

its formation should be avoided. Thus, the XRD results indicated


3.539
2h

that the catalysts calcined at 850 °C for 2 h and 900 °C for 5 h were
15
39

38
83
1
1
6


Elemental compositions, CaO concentrations and BET surface areas of the developed catalysts.

the best since they did not contain Ca2SiO4. Table 2 also lists the
CaO concentrations of the developed catalysts. The CaO concentra-
Calcined CaCO3 Uncalcined catalyst
after metal loading

tion of CaCO3 increased from 70% to 88 wt.% upon calcination at


800 °C. The uncalcined catalyst had a CaO concentration of only
56 wt.%. However, upon calcination under different conditions,
the CaO concentration significantly improved up to a maximum
0.485

of 83 wt.%. Calcination at 850 °C for 2 h and 900 °C for 5 h were


44

18
56
30

1
1
6

the best two conditions for maximum CaO concentrations. The


CaO concentrations obtained from the XRD analysis were in good
agreement with the Ca concentrations obtained from the EDS
at 800 °C

analysis.
9.63
43

51
88
6





3.1.5. Brunauer–Emmett–Teller (BET) surface area analysis


Catalyst type
Uncalcined

Table 2 also lists the BET surface areas of the catalysts devel-
oped under different calcination conditions. The surface area of
CaCO3

3.567 0.123

CaCO3 increased significantly from 0.123 to 9.63 m2 g1 upon cal-


16

34
50

70




cination at 800 °C. Likewise, the surface area significantly in-


creased from 0.485 m2 g1 for the uncalcined catalyst to a
ash
Fly

11
48

18

10

maximum of 6.399 m2 g1 for the calcined catalyst at 850 °C for


5
1

4

Mg

5 h. The catalysts calcined under different calcination tempera-


Ca
Al
Si
O

K
C

tures and durations displayed a wide range of surface areas from


CaO concentration

1.443 to 6.399 m2 g1. This indicated that the calcination condi-


BET surface area
composition

tions influenced the formation of pores within the catalysts. For


(m2 g1)
Elemental

the 2-h calcined catalysts, the surface areas increased as the tem-
(wt.%)

(wt.%)

perature increased up till 1000 °C when an unexpected drop in sur-


Table 2

face area occurred. Meanwhile, for the 5-h calcined catalysts, the
surface areas reduced as the temperature increased. The measured

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx 7

(a) (e)

Intensity (counts)

(b) (f)

(c) (g)

(d) (h)

Fig. 3. Powder XRD patterns of catalysts calcined at (a) 850 °C, (b) 900 °C, (c) 950 °C and (d) 1000 °C for 2 h; (e) 850 °C, (f) 900 °C, (g) 950 °C and (h) 1000 °C for 5 h
[characteristic peaks due to CaO (H), mullite (j) and dicalcium silicate (N)].

surface area trends are on the whole consistent with the SEM 950 and 1000 °C, respectively. The increase in calcination duration
observations of reduced morphological sizes with increased calci- was considered undesirable as it did not alter the biodiesel yield
nation temperatures for the 2-h cases. For the 5-h cases, as temper- significantly, but would increase the energy consumption. For the
ature increased, the surface areas reduced, but the morphological 2-h calcined catalysts, as the temperature increased from 850 to
sizes of the particles remained constant. The evidence seen here 1000 °C, the average yields decreased from 58.18% to 54.67%.
points to the longer calcination duration damaging the pores at Meanwhile, for the 5-h calcined catalysts, an increase in average
the catalyst surfaces. This resulted in the blockage of the pores yields from 51.77% to 62.89% was obtained as temperature in-
and subsequently, the reduction in surface areas although the mor- creased from 850 to 950 °C. This was followed by a decrease to
phological sizes remained the same. The catalysts with the lowest 57.35% at 1000 °C. The results suggested that the calcination
surface areas were obtained from calcinations at 1000 °C for both 2 temperature exerted a greater influence on the average yield than
and 5 h. This could also possibly be attributed to the blocking of the calcination duration. This is in line with Ngamcharussrivichai
pores within the ash supported CaO catalyst. An inversely propor- et al.’s [32] work in which it was reported that the increase in cal-
tional trend between surface area and Ca concentration is also ob- cination temperature exerted a much greater influence on the CO2
served for both the 2-h and 5-h calcined catalysts. It has been dissociation rate as compared to the calcination duration. Since the
reported that when CaO concentration increases, the mesopore dissociation of CO2 progresses from the outer surface into the par-
volume of the catalyst decreases which leads to lower surface area ticle, a CO2 film is formed at the surface leading to the recarbona-
[34]. tion of CaO to CaCO3. This recarbonation negatively affects the
catalytic performance of the catalyst.
3.2. Transesterification of CPO From Table 3, it can be seen that the catalyst calcined at 850 °C
for 2 h resulted in the highest biodiesel yield of 75.73%. This is con-
3.2.1. Optimisation of catalyst final calcination conditions sistent with the characterisation results which indicated that the
Table 3 lists the biodiesel yields from the transesterification catalyst calcined at 850 °C for 2 h displayed high CaO and Ca con-
reactions catalysed by waste fly ash supported CaO catalysts devel- centrations. Thus, the catalyst had high basicity despite having a
oped under different final calcination conditions. The catalysts cal- lower surface area. Additionally, this catalyst displayed a high ther-
cined for 2 h resulted in higher biodiesel yields at lower calcination mal stability as inferred from its TGA. Hence, it could be concluded
temperatures (850 and 900 °C). Conversely, the catalysts calcined that a calcination temperature of 850 °C was sufficient to produce
for 5 h had higher yields at higher calcination temperatures (950 an effective catalyst. This temperature also matched the decompo-
and 1000 °C). At lower temperatures, the increase in calcination sition temperature of CaCO3. Upon calcination to 850 °C, CaO is
duration from 2 to 5 h resulted in the reduction of the average formed, which has many defects in its crystal structure. These de-
yields (of all catalyst loadings) from 58.18% to 51.77%, and fects favour the formation of calcium methoxide, which is a surface
58.09% to 54.09% for 850 and 900 °C, respectively. In contrast, at intermediate in transesterification. Generally, for 2-h calcined cat-
higher temperatures, the increase in calcination duration improved alyst, when the calcination temperature exceeded 900 °C, the cat-
the average yields from 56.93% to 62.89%, and 54.67% to 57.35% for alytic activity gradually decreased despite the increase in BET

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
8 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

Table 3 sites available for methanol to form methoxide anion. However,


Biodiesel yields from transesterification reactions catalysed by developed catalysts. both yield and conversion decreased when the catalyst loading ex-
Calcination Calcination Catalyst loading Biodiesel ceeded 6 wt.%, down to 49.41% and 84.82%, respectively at 8 wt.%
temperature (°C) duration (h) (wt.%) yield (%) loading. This could possibly be due to the emulsion formed by
850 2 2 51.22 the high catalyst content leading to soap formation and reduced
4 55.18 biodiesel production. Furthermore, at high catalyst loading, the ef-
6 75.73 fect of mass transfer intensification becomes more dominant than
8 50.59
5 2 43.96
the effect of increased catalyst amount [36]. The high reaction mix-
4 47.64 ture viscosity would result in poor mass diffusion within the meth-
6 63.53 anol-oil-heterogeneous catalyst system [37].
8 51.94 Fig. 4(b) depicts the effects of methanol to oil molar ratio on
900 2 2 52.81 biodiesel yield and FAME conversion. Since transesterification is
4 58.38 a reversible reaction, sufficient methanol must be present during
6 70.59
the transesterification process to trigger and drive the reaction for-
8 50.59
5 2 45.77 ward. By increasing the methanol to oil molar ratio from 6:1 to 9:1,
4 46.82 the biodiesel yield and FAME conversion increased from 73.06% to
6 65.41 73.42%, and 82.41% to 94.64%, respectively. Further increment to
8 58.35 12:1 improved both the yield and conversion, whereby a maxi-
950 2 2 50.10 mum of 75.73% yield and 98.30% conversion were achieved. How-
4 55.27 ever, a reduction in yield from 75.73% to 54.12% was observed
6 63.53
8 58.82
when the molar ratio exceeded 12:1. Meanwhile, the conversion
5 2 51.34 dropped to almost constant at 97%. A methanol to oil molar ratio
4 71.76 of 12:1 was deemed to be the best ratio offering the highest yield.
6 73.18 Above this ratio, the higher methanol concentration leads to de-
8 55.29
creased yields due to the increased methanol–glycerol solubility
1000 2 2 49.62 which interferes with the glycerol separation [37]. The presence
4 71.29
of the polar hydroxyl group in methanol promotes emulsification
6 51.17
8 46.59 of the product [38].
5 2 54.41 The effects of reaction time on biodiesel yield and FAME conver-
4 59.76 sion can be seen in Fig. 4(c). At 1 and 2 h, experimental tests
6 61.89 showed that the conversions were very low at below 50% (data
8 53.34
not shown). This was caused by the slow reaction rate due to dis-
Transesterification conditions: Methanol to oil molar ratio 12:1; temperature 60 °C, persion and mixing between the methanol and oil. Generally,
time 3 h; stirring speed 700 rpm. increasing the reaction time promotes the transesterification pro-
cess towards completion. A maximum of FAME conversion of
98.30% was observed at the end of 3 h and thereafter it remained
surface area. The effects on the catalytic activity were not merely a approximately constant at 97.50%. The slight drop in conversion
result of the change in the catalyst surface area, but also of the could be partly associated to the formation of glycerol under longer
transformation within the crystal structure. The structure became duration [39]. In contrast, the biodiesel yield decreased when the
more intact with the rise in calcination temperature. The coordina- reaction time exceeded 3 h. This was because a longer duration en-
tion of Ca2+ and O2 on the catalyst surface changed which varied hanced the hydrolysis of esters (reversed transesterification)
the amount and strength of basic sites thereby significantly affect- resulting in the loss of esters as well as causing more fatty acids
ing the catalytic performance [13]. The findings obtained in this to form soap [40]. The results here indicated that the suitable time
work indicated that the overall catalytic activity does not depend required for achieving the highest amount of biodiesel yield was
solely upon its surface area or other factors such as basicity, 3 h.
porosity and morphology size, but the combined effects of these Fig. 4(d) illustrates the effects of reaction temperature on bio-
multiple attributes [35]. diesel yield and FAME conversion. In a methanol-oil-heteroge-
neous catalyst system, the reaction rate decreases because of
3.2.2. Effects of transesterification reaction parameters diffusional resistance between the liquid–liquid–solid phases.
To investigate the effects of transesterification reaction param- However, the reaction rate can be accelerated with higher reaction
eters on biodiesel yield and FAME conversion, the catalyst devel- temperatures. From Fig. 4(d), it could be observed that a low reac-
oped using a calcination temperature of 850 °C and duration of tion temperature of 45 °C was sufficient to achieve a high FAME
2 h was employed in a single factor parametric study. Fig. 4(a) conversion of 97.09%. As the reaction temperature increased from
shows the effects of catalyst loading on biodiesel yield and FAME 45 to 65 °C, the FAME conversion improved slightly from 97.09%
conversion. Both biodiesel yield and FAME conversion increased to 98.59% as higher temperatures increase reaction rates. However,
as the catalyst loading increased. The biodiesel yield increased the highest FAME conversion might not be necessarily accompa-
from 51.22% to 75.73% while the FAME conversion improved nied by the highest yield. A decreasing trend of yield was observed
slightly from 96.57% to 98.30% when the catalyst loading increased with increasing reaction temperature. At the highest temperature
from 2 to 6 wt.%. Thus, it could be seen that the developed waste of 65 °C, the highest FAME conversion of 98.59% was attained to-
fly ash supported CaO catalyst positively affected the yield and gether with the lowest yield of 73.53%. At this temperature, the
conversion due to its basicity. At low catalyst loading of 2 wt.%, boiling reaction mixture formed a large number of bubbles, which
although the FAME conversion exceeded the minimum ester con- inhibited the reaction especially on the three-phase interfaces [36].
tent requirement, the low biodiesel yield of 51.22% was undesir- Furthermore, higher temperatures can accelerate the side saponifi-
able. The maximum biodiesel yield reached 75.73% with a high cation of triglycerides [40]. The combination of these two effects
FAME conversion of 98.30% when 6 wt.% of catalyst loading was likely led to the decrease in the final biodiesel yield. Nevertheless,
used. This increment could be attributed to the increase in basic the differences in both yield and conversion are small at 1.5% and

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx 9

(a) (b)

(c) (d)

Fig. 4. Effects of (a) catalyst loading, (b) reaction temperature, (c) methanol to oil molar ratio and (d) reaction time on biodiesel yield and FAME conversion under nominal
parametric settings.
FAME conversion (%)

-ln (1-X)

(a) Reaction time (min) (b) Reaction time (min)

(c)
Fig. 5. (a) FAME conversion versus time at 6 wt.% catalyst loading, 12:1 methanol to oil molar ratio and 700 rpm stirring speed under different reaction temperatures, (b)
ln(1  X) versus t at different reaction temperatures and (c) Arrhenius plot of ln k versus 1/T  103.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
10 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

6.13%, respectively. This indicated that temperature was not a Table 4


dominant factor when the reaction time was long. A reaction with Rate constants k and correlation coefficients R2 for the transesterification of CPO at
different temperatures.
higher temperature can attain the maximum FAME conversion or
equilibrium state in a shorter time, but will remain constant there- Temperature (°C) Rate constant, k (min1) R2
after. With this in mind, a temperature of 45 °C was sufficient to 45 0.0041 0.944
achieve a yield of 79.76% and an FAME conversion of 97.09%. Addi- 50 0.0063 0.916
tional increase in reaction temperature above this was not neces- 55 0.007 0.936
60 0.0082 0.956
sary in the case of the current work considering the increase in 65 0.0118 0.942
energy consumption.

3.3. Kinetic modelling 3.4. Biodiesel physicochemical properties

Fig. 5(a) illustrates that FAME conversion increases as reaction The key biodiesel physicochemical properties are tabulated in
time increases. FAME conversion was slow in the initial stage. This Table 5. The measured ester and linolenic acid contents, kinematic
was followed by a more rapid increase and finally, the conversion viscosity, density, flash point, acid value, HHV and CN are within
reached equilibrium at 180 min. An increasing trend in the FAME the limits specified by the European Standard EN 14214 [24]. It
conversion with temperature was observed since higher tempera- is important that the properties meet the requirements set since
tures favoured the reaction. However, the highest FAME conver- they impact both the structure and performance of diesel engines.
sion obtained at 65 °C did not necessarily produce the highest For instance, to prevent coking on the injectors, kinematic viscosity
biodiesel yield. On the contrary, the lowest yield was obtained is a critical consideration. The measured kinematic viscosity at
due to saponification which occurred at high temperatures [41]. 40 °C was slightly below the upper limit of 6.0 mm2 s1. The mea-
By plotting ln(1  XME) versus t as shown in Fig. 5(b), strong sured density at 15 °C was slightly above the lower limit of
linear correlations were obtained which supported the hypothesis 860 kg m3. The flash point of the biodiesel was 144 °C. The high
of pseudo-first order reaction. Table 4 summarises the k values for flash point value ensures that the fuel is safe in terms of storage
the temperatures studied and the corresponding correlation coeffi- and transportation. The acid value was determined to be
cients (R2) obtained from the linear fitting. Fig. 5(c) shows that 0.42 mg KOH g1 which was below the upper limit of
string linearity (R2 = 0.95) could be observed between ln k and 1/ 0.5 mg KOH g1. An upper limit is imposed on acid value to avoid
T across the range of temperatures studied. The Ea determined corrosion related problems. The HHV was measured as
from the slope of the line was 42.56 kJ mol1, while the ko calcu- 38.1 MJ kg1 while the cetane number was predicted as 48.1. Both
lated from the intercept was 42192.6 min1. As such, the Arrhenius of these properties are indicators of the quality of the fuel.
equation for this heterogeneous catalysed reaction could be writ-
ten as: 3.5. Catalyst reusability

ln k ¼ 10:65  5:1181=T ð12Þ 3.5.1. Optimisation of catalyst regeneration method


Catalyst recycling is an important step as it reduces the overall
where k is the rate constant (min1) and T is the temperature (K). biodiesel production cost. Prior to recycling, the spent catalyst

Table 5
Physicochemical properties of produced biodiesel.

Physicochemical property Units Test method Limits Produced biodiesel


Kinematic viscosity mm2 s1; 40 °C ASTM D 445 1.9–6.0 5.9
Density kg m3; 15 °C ASTM D 1298 860–900 865.3
Flash point °C ASTM D 93 130 minimum 144
Acid value mg KOH g1 ASTM D 664 0.5 maximum 0.42
Higher heating value MJ kg1 ASTM D 4868 35 minimum 38.1
Cetane number – ASTM D613 47 minimum 48.1
Ester content % (mol mol1) EN 14103 96.5 minimum 97.1
Linolenic acid content % (mol mol1) EN 14103 12 maximum 0.18

Table 6
CaO concentrations and elemental compositions of fresh, spent and regenerated catalysts.

Fresh catalyst Spent catalyst Regenerated catalyst


Method 1a Method 2b Method 3c
CaO concentration (wt.%) 83 55 67 70 74
Elemental composition (wt.%) C 15 39 30 14 16
O 39 40 36 36 36
Mg 1 1 1 1 1
Al 1 – 2 1 1
Si 6 3 3 13 10
K – – – 3 2
Ca 38 17 28 32 34
a
Methanol washing.
b
Recalcination without methanol washing.
c
Methanol washing followed by recalcination.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx 11

needs to be regenerated as failure to do so would deteriorate the


catalytic performance [14]. Table 6 lists the XRD and EDS results
of the fresh, spent and regenerated catalysts for the three different
regeneration methods. The CaO concentration decreased from
83 wt.% to 55 wt.% while the Ca concentration decreased from
38 wt.% to 17 wt.% after the first usage. From Table 6, methanol
washing followed by recalcination 850 °C for 2 h was identified
as the best regeneration method in which the CaO and Ca concen-
trations were regenerated to 74 wt.% and 34 wt.%, respectively.
These concentration were only 9 wt.% and 4 wt.% less than the ori-
ginal CaO and Ca concentrations, respectively. Recalcining the cat-
alyst converted the CaCO3 back to active CaO by removing CO2 and
other impurities. The element Al concentration after recalcining
was also acceptable. The Si concentration (major element of ash)
meanwhile increased upon recalcination. Fig. 6. Biodiesel yields and FAME conversions in reusability study (6 wt.% catalyst
loading, 12:1 methanol to oil molar ratio, 45 °C temperature, 3 h time and 700 rpm
stirring speed).
3.5.2. Reusability cycles
Catalyst reusability was tested under the optimum transesteri-
fication conditions of 6 wt.% catalyst loading, 12:1 methanol to oil
molar ratio, 45 °C reaction temperature, 3 h reaction time and Table 7
700 rpm stirring speed. After each cycle, the spent catalyst was CaO and Ca concentrations and BET surface areas of fresh and recycled catalysts.
regenerated using methanol washing followed by recalcination at Catalyst CaO concentration Ca concentration BET surface
850 °C for 2 h. As shown in Fig. 6, the yield and conversion dropped (wt.%) (wt.%) area (m2 g1)
to 38.93% and 53.79%, respectively in the fourth cycle. Biodiesel Fresh 83 38 3.539
was not produced in the cycle after that. Both the yield and conver- 2nd cycle 74 34 3.530
sion obtained in the fourth cycle were unacceptably low. The 3rd cycle 66 28 3.489
regenerated catalyst was active for three consecutive cycles 4th cycle 60 15 3.435
5th cycle 56 10 3.391
whereby the minimum yield and conversion were 60.75% and
70.72%, respectively. The decrease in yield could be attributed to
catalyst deterioration due to poisoning by glycerol and soap pres-
ent in the reaction mixture [36]. Dissolution of Ca from CaO to the transesterification cycle, the CaO concentration decreased from
alcoholic phase also caused the deactivation of active sites [42]. 83 wt.% to 74 wt.% which suggested that dissolution of CaO parti-
The presence of calcium glyceroxide in the spent catalyst could cles from the catalyst surface occurred [34]. After the third cycle,
possibly be a factor in its reduced reusability as reported by Kouzu the CaO concentration drastically fell to 60 wt.%, which was consis-
et al. [43]. Calcium glyceroxide is formed through the combination tent with the low observed catalytic activity. The decreases in yield
of CaO and glycerol [43]. Due to the low catalytic performance in and conversion after each cycle were also supported by the
the subsequent cycles, the yield could possibly be increased with decreasing trend shown by the Ca concentration from 38 to
the use of higher catalyst loading and longer reaction time. 15 wt.%.
Although ideally a catalyst should be reused as many times as pos- The fresh fly ash based-catalyst had a surface area of
sible, the limited number of reusability cycles with this fly ash sup- 3.539 m2 g1. The recycled catalysts had approximately the same
ported CaO is compensated by the fact that palm oil mill fly ash is a surface areas as compared to the fresh one. From this, it could be
waste product. inferred that the catalytic performance of the recycled catalysts
was linked more to the basicity property than the surface area.
3.5.3. Characterisation of the recycled catalysts Fig. 7 illustrates the SEM micrographs of the catalysts captured be-
The CaO and Ca concentrations of the recycled catalysts were fore and after the transesterification cycle. The fresh catalyst had
determined using XRD and EDS, respectively. The BET surface area uniform distribution of miniature agglomerates with irregular
was also measured. Table 7 shows the CaO and Ca concentrations shapes. After the reaction, the spent catalyst appeared to be made
and the BET surface areas of the fresh and recycled catalysts. The up of more condensed mass. Mootabadi et al. [42] have suggested
fresh fly ash based-catalyst had a high CaO concentration of that physical changes occurred after reaction and new active sur-
83 wt.% and a Ca concentration of 38 wt.%. The CaO concentration face morphological sites were formed, which had significant effect
decreased considerably after each subsequent cycle. After the first on catalytic activity.

(a) (b) (c) (d)

50 µm 50 µm 50 µm 50 µm

Fig. 7. SEM images (2000 magnification) of catalysts from (a) 1st cycle, (b) 2nd cycle, (c) 3rd cycle and (d) 4th cycle.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061
12 W.W.S. Ho et al. / Energy Conversion and Management xxx (2014) xxx–xxx

4. Conclusions [13] Zhu H, Wu Z, Chen Y, Zhang P, Duan S, Liu X, et al. Preparation of biodiesel
catalysed by solid super base of calcium oxide and its refining process. Chin J
Catal 2006;27:391–6.
This study focussed on the improved development and evalua- [14] Boey P-L, Gaanty PM, Shafida AH. Biodiesel production via transesterification
tion of a palm oil mill fly ash supported CaO catalyst. This hetero- of palm olein using waste mud crab (Scylla serrata) shell as a heterogeneous
catalyst. Bioresour Technol 2009;100:6362–8.
geneous base catalyst was used in the transesterification of CPO
[15] Boey P-L, Maniam GP, Hamid SA, Ali DMH. Utilisation of waste cockle shell
with methanol to yield biodiesel. The optimum final calcination (Anadara granosa) in biodiesel production from palm olein: optimisation using
conditions with 45 wt.% loading of calcined CaCO3 were 850 °C response surface methodology. Fuel 2011;90:2353–8.
[16] Oliveira DA, Benelli P, Amante ER. A literature review on adding value to solid
and 2 h. Under these conditions, 75.73% of biodiesel yield was ob-
residues: egg shells. J Clean Prod 2013;46:42–7.
tained. It was determined that transesterification conditions of [17] Ho WWS, Ng HK, Gan S. Development and characterisation of novel
6 wt.% catalyst loading, 12:1 methanol to oil molar ratio, 45 °C heterogeneous palm oil mill boiler ash-based catalysts for biodiesel
reaction temperature, 3 h reaction time and 700 rpm stirring speed production. Bioresour Technol 2012;1255:158–64.
[18] Sim JH, Kamaruddin AH, Bhatia S. The feasibility study of crude palm oil
resulted in biodiesel yield and FAME conversion of 79.76% and transesterification at 30 °C operation. Bioresour Technol 2010;10:8948–54.
97.09%, respectively. The kinetic study suggested that the hetero- [19] Melero JA, Bautista LF, Morales G, Iglesias J, Sanchez-Vazquez R. Biodiesel
geneous catalysed transesterification was a pseudo-first order production from crude palm oil using sulfonic acid-modified mesostructured
catalysts. Chem Eng J 2010;161:323–31.
reaction. The Ea of the reaction was determined as 42.56 kJ mol1. [20] Chakraborty R, Bepari S, Banerjee A. Transesterification of soybean oil
Key biodiesel physicochemical properties including ester and lino- catalysed by fly ash and egg shell derived solid catalysts. Chem Eng J
lenic acid contents, kinematic viscosity, density, flash point, acid 2010;165:798–805.
[21] Hayyan A, Alam MdZ, Mirghani MES, Kabbashi NA, Nazashida NI, Siran YM,
value, HHV and cetane number were within the limits specified et al. Sludge palm oil as a renewable raw material for biodiesel production by
in EN 14214. The developed catalyst could also be reused feasibly two-step processes. Bioresour Technol 2010;101:7804–11.
up to three consecutive cycles after regeneration using methanol [22] Atapour M, Kariminia H-R. Characterisation and transesterification of Iranian
bitter almond oil for biodiesel production. Appl Energy 2011;88:2377–81.
washing followed by recalcination at 850 °C for 2 h. Overall, the
[23] Leung DYC, Guo Y. Transesterification of neat and used frying oil: optimisation
potential of this low-cost heterogeneous base catalyst has been for biodiesel production. Fuel Process Technol 2006;87:883–90.
demonstrated for transesterification applications. [24] EN 14214. Automotive fuels – Fatty Acid Methyl Esters (FAME) for diesel
engines – requirements and test methods; 2003.
[25] Lempers HEB, Sheldon RA. The stability of chromium in CrAPO-5, CrAPO-11,
and CrS-1 during liquid phase oxidations. J Catal 1998;175:62–9.
Acknowledgements [26] Chipera SJ, Bish DL. Fitting full X-ray diffraction patterns for quantitative
analysis: a method for readily quantifying crystalline and disordered phases.
This work was supported by the Ministry of High Education Adv Mater Phys Chem 2013;3:47–53.
[27] Issariyakul T, Dalai AK. Comparative kinetics of transesterification for biodiesel
(MOHE), Malaysia as well as the Faculty of Engineering at The Uni- production from palm oil and mustard oil. Can J Chem Eng 2012;90:342–50.
versity of Nottingham Malaysia Campus. The authors would also [28] EN 14103. Fat and oil derivatives – Fatty Acid Methyl Esters (FAME) –
like to express gratitude to Havys Oil Mill Sdn. Bhd. for providing determination of ester and linolenic acid methyl esters contents; 2003.
[29] Cheenkachorn K. Prediction properties of biodiesels using statistical models
the crude palm oil and Seri Ulu Langat Palm Oil Mill for the supply and artificial neural network. Sust Energy Environ 2014;3003:176–9.
of palm oil mill boiler ash. [30] Zhang L, Sheng B, Xin Z, Liu Q, Sun S. Kinetics of transesterification of palm oil
and dimethyl carbonate for biodiesel production at the catalysis of
heterogeneous base catalyst. Bioresour Technol 2010;101:8144–50.
[31] Birla A, Singh B, Upadhyay SN, Sharma YC. Kinetics studies of synthesis of
References biodiesel from waste frying oil using a heterogeneous catalyst derived from
snail shell. Bioresour Technol 2012;106:95–100.
[1] Sharma YC, Singh B, Korstad J. Latest developments on application of [32] Ngamcharussrivichai C, Nunthasanti P, Tanachai S, Bunyakiat. Biodiesel
heterogeneous basic catalysts for an efficient and eco friendly synthesis of production through transesterification over natural calciums. Fuel Process
biodiesel: a review. Fuel 2011;90:1309–24. Technol 2010;91:1409–15.
[2] Granados ML, Alonso DM, Alba-Rubio AC, Mariscal R, Ojeda M, Brettes P. [33] Rashidi NA, Mohamed M, Yusup S. A study of calcination and carbonation of
Transesterification of triglycerides by CaO: increase of the reaction rate by cockle shell. World Acad Sci Eng Technol 2011;60:818–23.
biodiesel addition. Energy Fuels 2009;23:2259–63. [34] Witoon T, Bumrungsalee S, Vathavanichkul P, Palisakun S, Saisriyoot M,
[3] Helwani Z, Othman MR, Aziz N, Fernando WJN, Kim J. Technologies for Faungnawakij K. Biodiesel production from transesterification of palm oil with
production of biodiesel focusing on green catalytic techniques: a review. Fuel methanol over CaO supported on bimodal meso-macroporous silica catalyst.
Process Technol 2009;90:1502–14. Bioresour Technol 2014;156:329–34.
[4] Boey P-L, Gaanty PM, Shafida AH. Performance of calcium oxide as a [35] Li E, Rudolph V. Transesterification of vegetable oil to biodiesel over MgO-
heterogeneous catalyst in biodiesel production: a review. Chem Eng J functionalised mesoporous catalysts. Energy Fuels 2008;22:145–9.
2011;168:15–22. [36] Liu X, Piao X, Wang Y, Zhu S. Calcium ethoxide as a solid base catalyst for the
[5] Kouzu M, Hidaka J-S. Transesterification of vegetable oil into biodiesel transesterification of soybean oil to biodiesel. Energy Fuels 2008;22:1313–7.
catalysed by CaO: a review. Fuel 2012;93:1–12. [37] Kotwal MS, Niphadkar PS, Deshpande SS, Bokade VV, Joshi PN.
[6] Dias JM, Alvim-Ferraz MCM, Almeida MF, Diaz JDM, Polo MS, Utrilla JR. Transesterification of sunflower oil catalysed by flyash-based solid catalysts.
Biodiesel production using calcium manganese oxide as catalyst and different Fuel 2009;88:1773–8.
raw materials. Energy Convers Manage 2013;65:647–53. [38] Viriya-empikul N, Krasae P, Nualpaeng W, Yoosuk B, Faungnawakil K. Biodiesel
[7] Kouzu M, Kasuno T, Tajika M, Sugimoto Y, Yamanaka S, Hidaka J. Calcium oxide production over Ca-based solid catalyst derived from industrial wastes. Fuel
as a solid base catalyst for transesterification of soybean oil and its application 2012;92:239–44.
to biodiesel production. Fuel 2008;87:2798–806. [39] Calcens J-M, Pouilloux Y, Barrault J. Selective etherification of glycerol to
[8] Singh Chouhan AP, Sarma AK. Modern heterogeneous catalysts for biodiesel polyglycerols over impregnated basic MCM-41 type mesoporous catalysts.
production: a comprehensive review. Renew Sust Energy Rev Appl Catal A 2002;227:181–90.
2011;12:4378–99. [40] Eevera T, Rajendran K, Saradha S. Biodiesel production process optimisation
[9] Watkins RS, Adam FL, Wilson K. Li–CaO catalysed tri-glyceride and characterisation to assess the suitability of the product for varied
transesterification for biodiesel applications. Green Chem 2004;4:335–40. environmental conditions. Renew Energy 2009;34:762–5.
[10] MacLeod CS, Harvey AP, Lee AF, Wilson K. Evaluation of the activity and [41] Leung DYC, Guo Y. Transesterification of neat and used frying oil: optimization
stability of alkali-doped metal oxide catalysts for application to an intensified for biodiesel production. Fuel Process Technol 2006;87:883–90.
method of biodiesel production. Chem Eng J 2008;135:63–70. [42] Mootabadi H, Salamatinia B, Bhatia S, Abdullah AZ. Ultrasonic-assisted
[11] Alonso DM, Mariscal R, Granados ML, Maireles-Torres P. Biodiesel preparation biodiesel production process from palm oil using alkaline earth metal oxides
using Li/CaO catalysts: activation process and homogeneous contribution. as the heterogeneous catalysts. Fuel 2010;89:1818–25.
Catal Today 2009;143:167–71. [43] Kouzu M, Kasuno T, Tajika M, Yamanaka S, Hidaka J. Active phase of calcium
[12] Ngamcharussrivichai C, Totarat P, Bunyakiat K. Ca and Zn mixed oxide as a oxide used as solid base catalyst for transesterification of soybean oil with
heterogeneous base catalyst for transesterification of palm kernel oil. Appl refluxing methanol. Appl Catal A 2008;334:357–65.
Catal A 2008;341:77–85.

Please cite this article in press as: Ho WWS et al. Evaluation of palm oil mill fly ash supported calcium oxide as a heterogeneous base catalyst in biodiesel
synthesis from crude palm oil. Energy Convers Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.03.061

You might also like