Вы находитесь на странице: 1из 5

Process Optimization for Biodiesel Production from Waste Cooking

Palm Oil (Elaeis guineensis) Using Response Surface Methodology


L. H. Chin, B. H. Hameed,* and A. L. Ahmad
School of Chemical Engineering, Engineering Campus, UniVersity of Science Malaysia,
14300 Nibong Tebal, Penang, Malaysia
ReceiVed September 19, 2008. ReVised Manuscript ReceiVed NoVember 8, 2008
A central composite rotatable design was used to study the effect of methanol to oil ratio, reaction time,
catalyst amount, and temperature on the transesterication of waste cooking palm oil using oil palm ash as a
catalyst. The reaction was carried out at 10 bar. All of the variables except reaction time signicantly affected
the biodiesel yield, amount of catalyst and reaction temperature being the most effective, followed by methanol
to oil ratio. Using response surface methodology, a quadratic polynomial equation was obtained for biodiesel
yield by multiple regression analysis. The optimum conditions for transesterication of waste cooking palm
oil to biodiesel were found as follows: amount of catalyst of 5.35 wt% (based on oil weight), temperature of
60 C, methanol to oil ratio of 18.0 and reaction time of 0.5 h. The predicted and experimental biodiesel
yields were found to be 60.07% (wt) and 71.74% (wt), respectively.
1. Introduction
The high demand for energy in the industrialized world and
the pollution problems caused by the use of fossil fuels has
made necessary the development of alternative renewable energy
sources. One of the particular alternatives considered is the use
of biodiesel (transesterication from vegetable oil). Biodiesel
is a well known alternative, renewable fuel which produces
fewer harmful emissions than conventional fossil-based diesel
fuel.
1
Moreover, due to its biodegradability and nontoxicity, the
production of biodiesel is considered to be an advantage to that
of fossil fuels. Its use leads to a decrease of the carbon dioxide,
sulfur dioxide, unburned hydrocarbon, and particulate matter
emissions generated in the combustion process.
2
However, in
spite of the favorable impact, the economic aspect of biodiesel
production is still a barrier, as the cost of biodiesel production
is highly dependent on the cost of feedstock, which affects the
cost of the nished product by up to 60-75%.
3
Currently,
partially or fully rened and edible-grade vegetable oils, such
as soybean, rapeseed, and sunower, are the predominant
feedstocks for biodiesel production,
4
which obviously results
in the high price of biodiesel. Therefore, exploring ways to
reduce the cost of the raw material is of great interest. In the
present work, waste cooking palm oil was chosen as the raw
material to produce biodiesel.
Currently, the main process for the synthesis of biodiesel is
the transesterication of vegetable oils using a strong base as
the homogeneous catalyst. However, this process presents some
disadvantages, as it requires the use of high amounts of catalyst
(which cannot be recovered), the production of different streams
which might be treated (neutralization step and wash step), and
the purication of glycerine to reuse it. These aspects also play
important roles in the economy of the process.
5
The use of heterogeneous catalysts is related to the develop-
ment of an environmentally benign process and the reduction
of the production cost.
6
Recently, several studies on the
transesterication of triglycerides have been conducted using
heterogeneous catalysts such as supported CaO,
7
calcium
ethoxide,
8
MgO-functionalized mesoporous catalyst,
9
KF/
Eu
2
O
3
,
10
MgO loaded with KOH,
11
and KF/Hydrotalcite.
12
The aim of this work was to study the performance of empty
fruit palm ash, a waste of the palm oil industry, as a catalyst in
the transesterication of waste cooking palm oil. Most of the
studies on the transesterication changed one separate factor at
a time. However, a reaction system simultaneously inuenced
by more than one factor can be poorly understood with the
change one separate factor at a time approach.
13
Therefore, the
experiments were performed according to central composite
design (CCD) and respond surface methodology (RSM) to
understand the relationship between the factor and yield to
biodiesel, and to determine the optimum conditions for produc-
tion of biodiesel.
2. Experimental Section
Waste cooking palm oil with a kinematic viscosity of 38.37
mm
2
s
-1
was collected from the canteen of the Engineering Campus,
* To whom correspondence should be addressed. Tel: +604-599 6422.
Fax: +604-594 1013. E-mail: chbassim@eng.usm.my.
(1) Issariyakul, T.; Kulkarni, M. G.; Meher, L. C.; Dalai, A. K.; Bakhshi,
N. N. Chem. Eng. J. 2008, 140, 7785.
(2) Antol n, G.; Tinaut, F. V.; Briceno, Y.; Castano, V.; Perez, C.;
Ram rez, A. I. Bioresour. Technol. 2002, 83, 111114.
(3) Cetinkaya, M.; Karaosmanoglu, F. Energy Fuels 2004, 18, 1888
1895.
(4) Haas, M. J. Fuel Process. Technol. 2005, 86, 10871096.
(5) Ramos, M. J.; Casas, A.; Rodr guez, L.; Romero, R.; P

; erez, A

.
Appl. Catal., A: General 2008, 346, 7985.
(6) Kim, H.-J.; Kang, B.-S.; Kim, M.-J.; Park, Y. M.; Kim, D.-K.; Lee,
J.-S.; Lee, K.-Y. Catal. Today 2004, 93-95, 315320.
(7) Yan, S.; Lu, H.; Liang, B. Energy Fuels 2008, 22, 646651.
(8) Liu, X.; Piao, X.; Wang, Y.; Zhu, S. Energy Fuels 2008, 22, 1313
1317.
(9) Li, E.; Rudolph, V. Energy Fuels 2008, 22, 145149.
(10) Sun, H.; Hu, K.; Lou, H.; Zheng, X. Energy Fuels 2008, 22, 2756
2760.
(11) Ilgen, O.; Akin, A. N., Energy Fuels 2008, doi:10.1021/ef800345u.
(12) Gao, L.; Xu, B.; Xiao, G.; Lv, J., Energy Fuels 2008, doi:10.1021/
ef800340w.
(13) Yuan, X.; Liu, J.; Zeng, G.; Shi, J.; Tong, J.; Huang, G. Renew.
Energy 2008, 33, 16781684.
Energy & Fuels 2009, 23, 10401044 1040
10.1021/ef8007954 CCC: $40.75 2009 American Chemical Society
Published on Web 01/06/2009
University of Science Malaysia, Penang. It was heated to 120 C
to remove excess water before use. Methanol (g99.9%, HPLC
Gradient grade) was purchased from Merck (Malaysia) and refer-
ence standards, such as methyl stearate (g99.5%), methyl palmitate
(g99.5%), methyl myristate (g99.5%), methyl oleate (g99.5%),
methyl linoleate (g99.5%) were employed, while methyl hepta-
decanoate (g99.5%) was used as an internal standard and was
purchased from Sigma-Aldrich (Malaysia) for gas chromatographic
analysis. N-hexane (g96%) was used as a solvent for GC analysis
and was purchased from Merck (Malaysia). All of the chemicals
used were of analytical reagent grade.
2.1. Catalyst. Oil palm ash was obtained from an oil palm mill
at Jawi, Penang, Malaysia. The precursor of the oil palm ash was
empty fruit bunches consisting of bers which were combusted at
800 C to generate energy for a boiler in the mill.
14
The oil palm
ash produced was observably coarse in nature. Large and unburned
residues were manually discarded. The ash was sieved and dried
in an oven overnight prior to use as a catalyst.
2.2. Experimental Design. The synthesis of biodiesel from palm
oil transesterication using oil palm ash as a catalyst was developed
and optimized using the Central Composite Design (CCD) and
Response Surface Methodology (RSM). CCD helps in investigating
linear, quadratic, cubic, and cross-product effects of the four reaction
condition variables on the biodiesel yield. The four independent
variables studied were reaction time, methanol to oil molar ratio,
temperature, and amount of catalyst. Table 1 lists the range and
levels of the four independent variables studied. Selection of the
levels was carried out on the basis of results obtained in a
preliminary study, considering limits for the experiment set-up and
working conditions for each chemical species. The value of R for
this CCD was xed at 2.
15
The complete design matrix of the
experiments employed and their results are given in Table 2. All
variables at the zero level constitute the center points and the
combination of each of the variables at either its lowest (-2.0)
level or highest (+2.0) level with the order variables at zero level
constitute the axial points. The experiment sequence was random-
ized to minimize the effects of the uncontrolled factors.
2.3. Statistical Analysis. The experimental data obtained by
following the above procedure were analyzed by the respond surface
methodology using the following second-order polynomial equation:
y )
0
+

i)1
n
B
i
x
i
+

i)1
n
B
ii
x
i
2
+

i<j)1
n
B
ij
x
i
x
j
(1)
where y is the response (biodiesel yield, wt%); x
i
and x
j
are the
coded independent variables and
0
,
i
,
ii
and
ij
are intercept,
linear, quadratic, and interaction constant coefcients, respectively.
Design Expert software version 6.06 (STATE-EASE Inc., Min-
neapolis, USA) was used for regression analysis and analysis of
variance (ANOVA). Response surfaces and contour plots were
developed using the tted quadratic polynomial equation obtained
from regression analysis, holding two of the independent variables
at a constant value corresponding to the stationary point and
changing the other two variables. Conrmatory experiment was
carried out to validate the equation, using combinations of
independent variables which were not part of the original experi-
mental design but within the experimental region.
2.4. Reaction Study. The transesterication of waste cooking
palm oil with methanol was carried out in a 300 mL stainless steel
batch reactor (PARR 4842 series reactor) consisting of a four-bladed
pitched turbine impeller, and a thermocouple. To ensure that most
of the reactants were in the liquid phase, the reactor was initially
pressurized at 10 bar. For the reaction startup, the catalyst was
initially charged into the reactant mixture at room temperature, after
which the reaction mixture was pressurized and heated to the desired
temperature. The reactor temperature was controlled with a
programmable PID temperature controller. The stirrer speed was
increased to maximum (300 rpm), which was sufcient to keep
the system uniform in temperature and suspension; this point was
taken as time zero for the reaction. At the end of the experiment,
the heater and stirrer were switched off and the reactor was cooled
to room temperature. After cooling, the catalyst was separated from
the product mixture by ltration. The product and byproduct were
collected and settled overnight to produce two phases: biodiesel at
top layer and glycerol at bottom layer.
The top layer product (biodiesel) was analyzed by gas chroma-
tography, Perkin-Elmer Clarus 500 with FID detector, equipped
with a polyethylene glycol column (Nukol, 15 m, 0.53 mm i.d.
and lm thickness of 0.5 m), at temperature from 110 to 190 C
and hold for 10 min, with a heating rate of 10 C/min. Helium was
used as a carrier gas. The injector and the detector temperatures
were set at 220 C and 250
c
C, respectively. Methyl heptadecanoate
was used as an internal standard. The biodiesel samples were diluted
with n-hexane and the injection volume was 1 L.
2.5. Biodiesel Yield. The amount of methyl esters in each sample
was quantied by comparing the signal of each methyl ester of the
gas chromatography of biodiesel products with the signal of each
of standards. The methyl esters yield in each experiment was
(14) Yin, C. Y.; Kadir, S. A. S. A.; Lim, Y. P.; Syed-Arifn, S. N.;
Zamzuri, Z. Fuel Process. Technol. 2008, 89, 693696.
(15) Montgomery, D. C., Design and Analysis of Experiments, 5th ed.;
John Willy and Sons: New York, 2001.
Table 1. Levels of the Transesterication Condition Variables
levels
variable coding units -2 -1 0 +1 +2
reaction time A hour 0.50 1.25 2.00 2.75 3.50
methanol to oil molar ratio B 6 9 12 15 18
temperature C C 60 95 130 165 200
amount of catalyst D wt% 1 4 7 10 13
Table 2. Experimental Design Matrix and Results
experimental variables
std
reaction
time, A (h)
methanol
to oil
molar ratio, B
temperature, C
(C)
amount of
catalyst, D
(wt%)
biodiesel
yield
(wt%)
1 1.25 9 95 4 45.21
2 2.75 9 95 4 46.66
3 1.25 15 95 4 51.25
4 2.75 15 95 4 49.58
5 1.25 9 165 4 47.87
6 2.75 9 165 4 49.90
7 1.25 15 165 4 41.57
8 2.75 15 165 4 41.86
9 1.25 9 95 10 42.26
10 2.75 9 95 10 50.27
11 1.25 15 95 10 45.47
12 2.75 15 95 10 44.75
13 1.25 9 165 10 42.95
14 2.75 9 165 10 45.54
15 1.25 15 165 10 23.14
16 2.75 15 165 10 27.73
17 0.50 12 130 7 45.84
18 3.50 12 130 7 50.65
19 2.00 6 130 7 42.75
20 2.00 18 130 7 43.32
21 2.00 12 60 7 52.96
22 2.00 12 200 7 41.03
23 2.00 12 130 1 41.05
24 2.00 12 130 13 25.03
25 2.00 12 130 7 45.43
26 2.00 12 130 7 50.35
27 2.00 12 130 7 49.65
28 2.00 12 130 7 47.71
29 2.00 12 130 7 42.41
30 2.00 12 130 7 48.24
Biodiesel Production from Waste Cooking Palm Oil Energy & Fuels, Vol. 23, 2009 1041
calculated from their content in the biodiesel as analyzed by gas
chromatography. The yield was dened as a ratio of the weight of
methyl esters, determined by gas chromatography, to the weight
of oil used.
16
2.6. Characterization of the Catalyst. Scanning electron
microscopy (SEM) analysis was carried out on the catalyst before
the transesterication reaction to study its surface morphologies
using eld-emission scanning electron microscope (FESEM) system
(Philips XL30S model). The sample was placed in a sample grid
and coated with gold-palladium for electron reection and
vacuumed before analysis. Energy dispersive X-ray (EDX) was used
to determine the elemental composition of oil palm ash by analyzing
the microscopic image under EDX instrument (Philips XL30S
model). Fourier transform infrared (FTIR) analysis was applied on
the same catalyst to determine the surface functional groups using
FTIR spectroscopy (FTIR-2000, Perkin-Elmer). The spectra were
recorded from 4000 to 400 cm
-1
.
3. Results and Discussion
3.1. Response Surface Methodology (RSM). A central
composite design (CCD) was used to develop a correlation
between the transesterication condition variables to the biodie-
sel yield. The complete design matrix and biodiesel yield at
various transesterication condition variables are listed in Table
2. The biodiesel yields obtained were in the range from 23.14
to 52.96 wt%. Runs 25 to 30 at the center point of the design
were used to determine the experimental error. The normal plot
of residuals for biodiesel yield (Figure 1) was normally
distributed and showed no deviation of the variance. Besides,
the model was also tested for any transformation that could have
been applied, but the Box-Cox plot did not suggest any
transformation for the response.
3.2. Regression Analysis. A regression analysis was per-
formed to t the response function and predict the outcome of
biodiesel yield with a simple equation. The model is expressed
by Eq. (2) which takes their coded value.
Biodiesel yield )47.68 +1.09 A-1.84 B-3.28 C-
3.49 D-1.00 B
2
-3.50 D
2
-3.66 B C-
2.16 B D-1.99 C D (2)
The summary of the analysis of variance (ANOVA) result is
shown in Table 3. The regressors or term incorporated in the
model are those statistically tested to be signicant. The Prob
> F value indicates that probability equals the proportion of
the area under the curve of the F-distribution that lies beyond
the observer F value. The small probability values called for the
rejection of the null hypothesis, in other words, the particular
term signicantly affected the measured response of the system.
In these cases, the Prob > F less than 0.05 indicated the
particular term was statistically signicant. The analysis con-
rmed the signicant terms of B, C, D, D
2
, BC, BD, and CD
for biodiesel yield. The term A was found not to be signicant
and null hypothesis in the biodiesel yield but was included into
the analysis for the sake of maintaining the hierarchical structure
of the model terms. The coefcient of determination, R
2
for
the model was 90.56%. This indicates that only 9.44% of the
total variability was not explained by the regressors in the model.
The high R
2
value species that the model obtained will be able
to give a convincingly good estimate of response of the system
in the range studied. The lack of t test, which is not signicant
for the model developed, shows that the model satisfactorily
ts the data. Figure 2 shows the predicted values versus actual
values for biodiesel yield. As can be seen, the predicted values
obtained were quite close to the experimental values, indicating
that the model developed was successful in capturing the
correlation between the transesterication condition variables
to the biodiesel yield.
3.3. Model Analysis. The result of regression analysis
seemed to suggest that biodiesel yield was only affected by the
main factor of methanol to oil molar ratio (B), temperature (C),
amount of catalyst (D), and their respective higher-order term
(D
2
). Signicant interactions terms were found to exist between
the main factors (BC, BD, and CD). Figure 3 shows the three-
dimensional response surface that was constructed to show the
effects of the transesterication condition variables (methanol
(16) Jitputti, J.; Kitiyanan, B.; Rangsunvigit, P.; Bunyakiat, K.; Attana-
tho, L.; Jenvanitpanjakul, P. Chem. Eng. J. 2006, 116, 6166.
Figure 1. Normal plot of residual for biodiesel yield.
Table 3. ANOVA for Model Regression
source SS
a
d.f.
a
mean
square F-value
probability
> F
model 1373.35 9 152.59 21.31 <0.0001 signicant
A 28.58 1 28.58 3.99 0.0595
B 81.29 1 81.29 11.35 0.0030
C 258.40 1 258.40 36.09 <0.0001
D 292.81 1 292.81 40.89 <0.0001
B
2
28.56 1 28.56 3.99 0.0596
D
2
348.58 1 348.58 48.68 <0.0001
BC 214.70 1 214.70 29.98 <0.0001
BD 74.61 1 74.61 10.42 0.0042
CD 63.56 1 63.56 8.88 0.0074
residual 143.21 20 7.16
lack of t 99.93 15 6.66 0.77 0.6842 not signicant
pure error 43.29 5 8.66
a
SS, sum of squares; d.f., degree of freedom.
Figure 2. Predicted vs actual biodiesel yield.
1042 Energy & Fuels, Vol. 23, 2009 Chin et al.
to oil molar ratio and reaction temperature) on biodiesel yield.
The reaction time and amount of catalyst were xed at zero
level. As can be seen from Figures 3 and 4, biodiesel yield
increases with an increase in the methanol to oil molar ratio at
low temperature and low amount of catalyst. The excess
methanol can promote the transesterication reaction forward
and also extract products, such as glycerin and methyl esters,
from the system to renew the surface of the catalyst.
7
However,
biodiesel yield decreases at higher reaction temperatures above
130 C and at higher amount of catalyst with increased in
methanol to oil molar ratio. This may be due to the large amount
of methanol diluting the oil and reducing the biodiesel yield.
12
The reaction temperature can inuence the reaction rate and
the biodiesel yield because the intrinsic rate constants are strong
functions of temperature. The study of the effect of temperature
is very important for a catalyzed reaction.
17
From Figures 3
and 5, the results show that the biodiesel yield increased with
an increase of reaction temperature at low amount of catalyst
and a low methanol to oil molar ratio. Being an equilibrium
reaction, the equilibrium constant is inuenced by temperature
and pressure. In addition, in our experiments, which were carried
out at higher pressure, both factors affected the equilibrium
constant. Therefore, as the temperature increased, the biodiesel
yield increased. Moreover, because of the solid catalyst used
in this reaction, the mass transfer effect should be considered.
A high temperature is a benet to the mass transfer.
12
Neverthe-
less, the biodiesel yield dropped obviously at high temperature
for higher methanol to oil molar ratio and higher amounts of
catalyst. This may possibly be due to presence of solid catalyst,
the reaction mixture constitutes a three-phase system, oil-
methanol-catalyst, in which the reaction would be slowed down
because of the diffusion resistance between different phases.
18
This might be the reason for lower biodiesel yield at higher
methanol to oil molar ratio and higher amounts of catalyst.
In the case of homogeneous catalysts, it has been revealed
that the amount of catalyst has a strong inuence on the
conversion of vegetable oil to ester.
19
Thus, the effect of the oil
palm ash as a catalyst on the transesterication of waste cooking
palm oil was studied. From Figures 4 and 5, the biodiesel yield
increased with an increase in the amount of catalyst. This result
indicates that with the addition of more catalyst, there was a
faster rate at which the reaction equilibrium was reached because
of the increase in the total number of available active catalytic
sites for the reaction.
17
Nevertheless, biodiesel yield decreases
after the amount of catalyst exceeded 7 wt.% from waste
cooking oil used. This may be due to the presence of the solid
catalyst, the reaction rate is determined by surface reaction and
mass transfer.
17
Higher catalyst dosage may make the reactant
mixture more viscous, which will increase the mass transfer
resistance in the multiphase system.
6
Besides, the decrease in
biodiesel yield was also observed by Li and Xie
20
which is most
likely due to a mixing problem involving reactants and the solid
catalyst.
3.4. Optimization of Biodiesel Yield. The CCD was able
to function as an optimal design for the desired response of the
system based on the model obtained and the input criteria. The
optimization of biodiesel yield was carried out based on all
transesterication variables, which were in the range of experi-
mental runs. The software predicted that optimized conditions
for biodiesel yield were obtained when the reaction time,
methanol to oil molar ratio, temperature, and amount of catalyst
were at 0.5 h, 18.0 molar ratio, 60 C and 5.35 wt% of catalyst,
(17) Liu, X.; Piao, X.; Wang, Y.; Zhu, S.; He, H. Fuel 2008, 87, 1076
1082.
(18) Liu, X.; He, H.; Wang, Y.; Zhu, S.; Piao, X. Fuel 2008, 87, 216
221.
(19) Arzamendi, G.; Campo, I.; Arguinarena, E.; Sanchez, M.; Montes,
M.; Gandia, L. M. Chem. Eng. J. 2007, 134, 123130.
(20) Li, H.; Xie, W. J. Am. Oil Chem. Soc. 2008, 85, 655662.
Figure 3. Three-dimensional response surface plot of biodiesel yield
(effect of temperature and methanol to oil molar ratio, reaction time )
2 h, amount of catalyst ) 7 wt%).
Figure 4. Three-dimensional response surface plot of biodiesel yield
(effect of amount of catalyst and methanol to oil molar ratio, reaction
time ) 2 h, temperature ) 130 C).
Figure 5. Three-dimensional response surface plot of biodiesel yield
(effect of amount of catalyst and temperature, reaction time ) 2 h,
methanol to oil molar ratio ) 12).
Biodiesel Production from Waste Cooking Palm Oil Energy & Fuels, Vol. 23, 2009 1043
respectively, with predicted biodiesel yield of 60.07% (wt). The
experimental biodiesel yield was 71.74% (wt). This means that
the experimental value obtained was in good agreement with
the value calculated from the model.
3.5. Characterization of Oil Palm Ash Catalyst. Figure 6
shows the SEM image of the oil palm ash, which illustrates the
spongy and porous nature of the ash particles. The porous nature
of oil palm ash was also reported by Yin et al.
21
and
Tangchirapat et al.
22
Table 4 shows the elemental composition
of oil palm ash used in this study. A signicant observation is
that oil palm ash contains high weight percentage of potassium,
while aluminum, zinc, and magnesium weight percentages are
comparatively low. Besides, due to the high weight percentage
of oxygen, it can be postulated that potassium, magnesium,
silicone, zinc, and aluminum may exist for the most part in oxide
form. Furthermore, it was found that the K
2
O was the cause of
the high catalytic activity and basicity of the catalyst.
23-25
Figure 7 shows the FTIR transmission spectrum of oil palm
ash. A prominent peak observed at about 3390 cm
-1
is assigned
to -OH band. The existence of phenols and alcohol groups are
supported by the presence of the bands at 1400-1300 cm
-1
,
attributed to deformation (OH), 1100-1000 cm
-1
(stretching
(C-O)), and 680-620 cm
-1
(out of plane bending (O-H)).
4. Conclusions
Oil palm ash, a waste from the oil palm industry, was found
to be suitable catalyst for the transesterication of waste cooking
palm oil to biodiesel. A central composite design was conducted
to study the effects of methanol to oil ratio, reaction time,
catalyst amount, and temperature on the transesterication of
waste cooking palm oil. The predicted and experimental
biodiesel yields were found to be 60.07% (wt) and 71.74% (wt),
respectively.
Acknowledgment. The authors acknowledge the research grant
provided by the Ministry of Science, Technology, and Innovation
(MOSTI), Malaysia under Science Fund grant (Project No. 03-01-
05-SF0207), that resulted in this work.
EF8007954
(21) Yin, C. Y.; Wan Ali, W. S.; Lim, Y. P. J. Hazard. Mater. 2008,
150, 413418.
(22) Tangchirapat, W.; Saeting, T.; Jaturapitakkul, C.; Kiattikomol, K.;
Siripanichgorn, A. Waste Manage. 2007, 27, 8188.
(23) Xie, W.; Li, H. J. Mol. Catal. A: Chem. 2006, 255, 19.
(24) Xie, W.; Peng, H.; Chen, L. Appl. Catal., A: General 2006, 300,
6774.
(25) Noiroj, K.; Intarapong, P.; Luengnaruemitchai, A.; Jai-In, S., Renew.
Energy 2008, doi:10.1016/j.renene.2008.06.015.
Figure 6. SEM image of oil palm ash.
Figure 7. FTIR transmission spectrum of oil palm ash.
Table 4. Elemental Compositions of Oil Palm Ash Used in This
Study by EDX
elements weight (%)
potassium (K) 40.59
oxygen (O) 29.36
carbon (C) 14.56
silicone (Si) 2.63
magnesium (Mg) 0.76
phophorus (P) 0.73
aluminum (Al) 0.50
zinc (Zn) 0.33
clorin (Cl) 7.07
1044 Energy & Fuels, Vol. 23, 2009 Chin et al.

Вам также может понравиться