Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Bioresource Technology 110 (2012) 628–636

Contents lists available at SciVerse ScienceDirect

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

Gasification of oil palm empty fruit bunches: A characterization and kinetic study
M.A.A. Mohammed a, A. Salmiaton a,b,⇑, W.A.K.G. Wan Azlina a,b, M.S. Mohamad Amran a,b
a
Department of Chemical & Environmental Engineering, Faculty of Engineering, University Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
b
Institute of Advanced Technology, University Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia

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

Article history: Empty fruit bunches (EFBs), a waste material from the palm oil industry, were subjected to pyrolysis and
Received 19 July 2011 gasification. A high content of volatiles (>82%) increased the reactivity of EFBs, and more than 90%
Received in revised form 5 January 2012 decomposed at 700 °C; however, a high content of moisture (>50%) and oxygen (>45%) resulted in a
Accepted 13 January 2012
low calorific value. Thermogravimetric analysis demonstrated that the higher the heating rate and the
Available online 28 January 2012
smaller the particle size, the higher the peak and final reaction temperatures. The least squares estima-
tion for a first-order reaction model was used to study the degradation kinetics. The values of activation
Keywords:
energy increased from 61.14 to 73.76 and from 40.06 to 47.99 kJ/mol when the EFB particle size increased
Empty fruit bunch
Gasification
from 0.3 to 1.0 mm for holocellulose and lignin degradation stages, respectively. The fuel characteristics
Kinetics of EFB are comparable to those of other biomasses and EFB can be considered a good candidate for
gasification.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction changes that occur in the chemical structure of EFBs, which will af-
fect their performance when exposed to elevated temperatures
The oil palm (Elaeis guineensis) is the dominant agricultural crop during thermochemical conversion (Omar et al., 2011).
in many countries. Malaysia alone produces approximately 47% of Thermogravimetric analysis (TGA) is one of the techniques used
the world’s supply of palm oil and can be considered the world’s for examining the decomposition of solids, and it has been widely
largest producer and exporter of palm oil (Mohammed et al., used to investigate the thermal decomposition behavior of differ-
2011). Currently, more than 3.88 million hectares of land are under ent types of biomasses (Shen et al., 2009). Interpretation of the
oil palm cultivation and more than 368 palm mills operate in experimental data can provide information regarding biomass
Malaysia (Idris et al., 2010). In addition to palm oil, Malaysia composition, reaction order, and corresponding kinetic constants
generates huge quantities of oil palm biomass wastes, including (Quan et al., 2009). For an adequate design of gasification equip-
trunks, fronds, fibers, shells and empty fruit bunches (EFBs). Each ment, knowing the activation energy and the rate of thermal
fresh fruit bunch (FFB) processed in an oil mill generates 14% decomposition of biomasses is necessary, which relies on kinetic
fibers, 7% shells, and between 20% and 25% EFB, which is the ligno- studies of the concerned biomasses during the heating process.
cellulosic fibrous medium left behind after oil extraction (Basiron, Many models have been developed to establish the required
2007). Although a part of the EFB is either burnt in incinerators to kinetic parameters (i.e. activation energy and pre-exponential fac-
produce fertilizers or used as a solid fuel in boilers to generate tor) of biomass pyrolysis and gasification such as the distributed
steam and electricity in the palm oil mills, large quantities of EFBs activation energy model (Sonobe and Worasuwannarak, 2008),
have no specific use and need to be constructively utilized (Lahijani least-squares estimation method for a first-order reaction model
and Zainal, 2011). Compared to burning, incineration, and other (Orfao et al., 1999; Li et al., 2008) and Kissinger’s model-free kinet-
methods, pyrolysis and gasification can be considered as viable ics equation (Harris and Zhong, 2008). Among all these models, the
technologies for managing EFBs and simultaneously producing least-squares estimation method for a first-order reaction model
valuable liquid and gaseous biofuels. seems more applicable to the investigation of the behavior pat-
Biomass characteristics such as moisture, ash, volatile matter, terns of various biomass samples, particularly for three indepen-
inorganic elements, structural constituents, calorific value, particle dent parallel reactions involving the cellulose, hemicellulose and
size, and density are of great importance in understanding the lignin contents of biomass (Li et al., 2008).
In the present study, the chemical, physical, fuel-oriented and
thermal characteristics of EFBs are presented. Additionally, the ther-
⇑ Corresponding author at: Department of Chemical & Environmental Engineer-
mal decomposition behaviors of different EFB particle sizes were
ing, Faculty of Engineering, University Putra Malaysia, 43400 UPM Serdang,
Selangor, Malaysia. Tel.: +60 3 89466297; fax: +60 3 86567120. thermogravimetrically analyzed. Furthermore, the kinetic behavior
E-mail address: mie@eng.upm.edu.my (A. Salmiaton). of EFB is reported using the least-squares estimation model.

0960-8524/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2012.01.056
M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636 629

2. Methods combustion, with proper allowance for thermochemical and heat


transfer corrections. The bomb calorimeter was calibrated by
2.1. Empty fruit bunch combusting a known mass, m, of standard benzoic acid which
has a known heat of combustion of 26.453 kJ/g. Three benzoic acid
EFB samples were collected from the Seri Ulu Langat Palm Oil samples of size around one gram were tested. The calibration fac-
Mill, Dengkil, Selangor, Malaysia. The EFB is the lignocellulosic tor for the bomb calorimeter, C, was calculated from the benzoic
material remaining as a by-product of the industrial process after acid calibrations as: C = (m26.453 kJ/g)/(Tmax  To).
removal of the nuts. The samples had more than 50 wt.% moisture,
and were in the form of whole bunches. The EFBs were manually 2.2.4. Apparent density
chopped into small pieces that could be fed into a shredder. Subse- The apparent density of the EFB samples was determined using
quently, the EFB was dried at 100 ± 3 °C for 4 h. A grinder (FRITSCH, a gas pycnometer (model Micromeritics, AccuPyc II 1340) with he-
Germany) with a screen size of 1.0 mm was used to obtain the lium as purging gas. For calibration, the pycnometer was filled with
feedstock less than 1.0 mm in size. A motorized sieve shaker distilled water essentially at room temperature and the mass of the
(AIMIL, India) was used to separate the milled EFB into fractions pycnometer and water determined. A calibration table was estab-
of different particle sizes as illustrated in Table 1. lished from the mass determined at the observed temperature.

2.2. Characterization of empty fruit bunch 2.2.5. Analysis of lignocellulosic content


Dried, ground EFB (1.0 g, W0) was placed in a crucible, weighed
2.2.1. Proximate analysis (W1), and transferred to the boiling position in a hot extraction unit
The moisture content (wet basis) of the as-received EFB samples (Fibertec System, USA). One-hundred mL of preheated neutral
was determined using a standard test method (ASTM E1756-01) detergent solution (30 g Na-lauryl sulfate, 18.61 g EDTA, 6.81 g
whereas, the contents of moisture (dry basis), volatile matter, fixed Borax, 4.56 g Na2HPO4 and 10 mL 2-ethoxy ethanol were dissolved
carbon and ash were determined using a thermogravimetric ana- in 1 L distilled water) was added to the sample and allowed to boil
lyzer (model Mettler Toledo, TGA/SDTA851, USA). Approximately for 1 h. The crucible was moved from the hot extraction unit to a
10 mg of air-dried EFB powder was placed in an alumina crucible cold extraction unit (Fibertec System, USA), washed with acetone,
and inserted into the TGA furnace and weighed. To estimate the and dried. The crucible was placed in an oven at 105 ± 3 °C for a
moisture content (dry basis) and volatile matter, the EFB sample minimum of 3 h, removed from the oven, allowed to cool to room
was continuously heated from room temperature to a temperature temperature in a desiccator and weighed (W2). The neutral deter-
of 600 °C at a rate of 10 °C/min under N2 flow at the rate of 10 mL/ gent fiber (NDF) content was calculated using the following
min. Then, the furnace temperature was further increased to equation:
 
900 °C at 10 °C/min under 10 mL/min of air flow to estimate the w1 —w2
fixed carbon content. The residue was considered to represent
NDF ð%Þ ¼  100 ð1Þ
w0
the ash content.
Acid detergent fiber (ADF) content was determined by following
the same procedure for NDF except a different detergent solution
2.2.2. Ultimate analysis
was used i.e. acid detergent solution (20 g cetyl trimethylammoni-
The percentages of carbon, hydrogen, nitrogen, sulfur and oxy-
um bromide dissolved in 700 mL distilled water. 27.56 mL of 96.7%
gen (by difference) of the EFB sample were determined after com-
sulfuric acid was then added to the solution and the topped up to
plete combustion of the sample using a CHNS/O Analyzer (model
1 L with distilled water). The ADF content was then calculated
LECO TruSpec CHN, USA) following the ASTM D-5291 method.
using the following equation:
Approximately 5.0 mg of grained EFB samples were placed in a  
tin capsule and the capsule was crimped. This capsule, together w1 —w2
ADF ð%Þ ¼  100 ð2Þ
with capsules serving as a blank and capsules containing a stan- w0
dard (sulfamethazine) were placed into the autosampler of the
For acid detergent lignin (ADL) analysis, a crucible with a
CHNS/O analyzer. The analyzer oxidation temperature was set at
sample generated during the ADF analysis process was placed in
1000 °C. Helium, oxygen and nitrogen were used as carrier com-
a 50-mL beaker, covered with 72% H2SO4 (at 15 °C) and stirred with
bustion, autosampler injection gases, respectively.
a glass rod for a minimum of 3 h. The crucible content was filtered
through a filter paper under vacuum and washed with hot water.
2.2.3. Heating value The filtrate was placed in an oven at 105 ± 3 °C for a minimum of
The heating value of EFB was determined by burning a weighed 3 h, removed from the oven, allowed to cool to room temperature
sample in an adiabatic oxygen-bomb calorimeter (model Parr in a desiccator and weighed (W3). The content of the crucible was
1341, USA) following the ASTM D240 method. EFB (0.5 g) was ignited in a 500 °C furnace for 2 h, placed in a 100 °C oven for 1 h
placed in a metal cup connected with a fuse wire; the cup was and transferred to a desiccator, cooled, and weighed (W4). The
placed inside the bomb calorimeter and charged with oxygen to ADL content was calculated using the following equation:
approximately 200 psi. The heat of combustion was computed  
from the temperature observations before, during, and after w3 —w4
ADL ð%Þ ¼  100 ð3Þ
w0

Table 1 The percentages of cellulose, hemicellulose, and lignin were


Particle size distribution of EFB. then computed as follows:
Feed particle size (mm) Mass fraction Cellulose ð%Þ ¼ ADF  ADL ð4Þ
Less than 0.3 18
0.3–0.5 60 Hemicellulose ð%Þ ¼ NDF  ADF ð5Þ
0.5–1.0 14
More than 1.0 8
Lignin ð%Þ ¼ ADL ð6Þ
630 M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636

2.2.6. Analysis of inorganic element X


3 X
3 X
3

The inorganic materials in EFB ash, such as potassium, calcium, ðW o  WÞ ¼ W i0  W ¼ ðW i0  W i Þ ð12Þ


i¼1 i¼1 i¼1
magnesium, phosphorus and sodium, were estimated with an en-
ergy dispersive X-ray fluorescence (XRF) spectrometer (EDX-HS substituting in Eq. (10), a new expression can be obtained as
series model, Shimadzu-720, Japan). follows:
X
3
2.3. Thermal decomposition and kinetics analysis ðwo  wÞ ¼ ðwi0  wi1 Þai ð13Þ
i¼1
A series of non-isothermal experiments were conducted for TGA
Generally,
and differential thermal analysis (DTG) using a simultaneous TGA–
DTA thermogravimetric analyzer (model Mettler Toledo, TGA/ ðwo  wÞ ¼ ðwo  w1 Þa ð14Þ
SDTA851, USA). Approximately 10 mg of air-dried EFB powder
and, by comparing Eqs. (13) and (14), a new term for a can be
was placed in an alumina crucible, placed inside a furnace and pro-
obtained as follows:
grammed for continuous heating from room temperature to
1000 °C at predetermined heating rates of 10, 20, and 30 °C/min X3
ðwi0  wi1 Þ
under airflow at 100 mL/min. The data were recorded by a data a¼ ð15Þ
i¼1
ðwo  w1 Þ
logging system, which provided listings of sample weights and
temperatures with time. Indium was used to calibrate the temper- or
ature reading and the instrument was weight calibrated according
X
3
to the manufacturer’s instructions. a¼ Z i ai ð16Þ
Kinetic parameters for pyrolysis and gasification reactions were i¼1
calculated using least-squares estimation for a first order reaction
where Z i ¼ ðw i0 wi1 Þ
ðwo w1 Þ
. The unknown parameters of the model such as
model. The following assumptions were considered: (1) the bio-
Ai, Ei, and zi are determined by the least-squares estimation method.
mass consists of moisture, hemicellulose, cellulose and lignin; (2)
The minimized summation error is illustrated in the following
the biomass contents reacted independently; (3) the combustion
equation (Orfao et al., 1999):
process occurs as a parallel reaction; (4) the cellulose and hemicel-
lulose DTG peaks that overlap with each other to represent a new N h
X i2
component, known as holocellulose. In this model the rate of the S¼ ðaj Þexp  ðaj Þmod ð17Þ
j¼1
reaction for the decomposition of a solid depends on the tempera-
ture and the amount of substance; thus, for a single reaction, a Here, the subscript j refers to the TGA data points collected and the
simple rate expression can be written as follows (Lapuerta et al., superscripts exp and mod are the experimental and modeling
2004): results, respectively. The average deviation is defined as:
qffiffiffi
S
da N
kðTÞf ðaÞ ð7Þ Deviation ð%Þ ¼ 100 ð18Þ
dt amax
The variable a is the degree of degradation and k is the rate con- where amax equals 100, and N is the number of data points. The
stant. According to the Arrhenius expression: deviation can be simplified as follows:
rffiffiffiffi
  S
E Deviation ð%Þ ¼ ð19Þ
k ¼ A exp  ð8Þ N
RT

where A is the pre-exponential factor of the reaction, E is the activa- 3. Results and discussion
tion energy, R is the gas constant and T is the temperature. Thus, for
an independent parallel reaction, Eq. (7) can be rewritten as: 3.1. EFB characterizations
 
dai Ei 3.1.1. Proximate and ultimate analyses
¼ Ai exp  f ðai Þ ð9Þ
dt RT The results of the proximate and ultimate analyses are shown in
Table 2. A comparison of the results with those from previous re-
Here, the variable ai is the degree of degradation in the ith stage as search on EFB, shell, fiber, beech wood and bituminous coal is also
presented below: presented. The moisture content (wet basis) of as-received EFB
ðwi0  wi Þ samples was more than 50%, and therefore a drying process was
ai ¼ ð10Þ needed before grinding the EFB sample into powder feedstock of
ðwi0w1 Þ
different particle sizes suitable for the gasifying system. In the case
where wi is the mass of the sample in the ith stage, and the sub- of a fluidized bed gasifier, high moisture biomass causes feeding
scripts 0 and 1 refer to the initial and residual amounts, and fluidization problems. Furthermore, wet biomass fuels require
respectively. a longer residence time for drying before gasification. In addition,
The function f (ai) is typically dependent on the particular the overall gasification efficiency decreases as the moisture con-
decomposition mechanism, and the most used expression for it is tent of the fuel increases due to reduction in the temperature of
represented below: the oxidation zone, resulting in incomplete conversion of the
hydrocarbons released from the pyrolysis zone. However, some
f ðai Þ ¼ ð1  ai Þni ð11Þ
level of moisture is advantageous to gasification reactions such
where (1  ai) represents the remaining portion of the convertible as the water gas shift reaction and the steam reforming reaction,
part of the biomass and n is the reaction order. As this EFB sample resulting in high syngas products (Kataki and Konwer, 2001).
consists of three-pseudocomponents, the decomposition rate can be As shown in Fig. 1, a low moisture content (dry basis) of EFB
expressed as follows: (5.18%), high percentage of volatile matter (82.58%), low amounts
M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636 631

Table 2
Properties of empty fruit bunch (EFBs), solid waste biomasses and coal.

Properties EFB (this study) EFB (literature) Oil palm shelld Oil palm fibere Beech woodf g
Coalf
Proximate analysis (wt.%)
Moisture (wb) 55.6 67a 30 35 –
Moisture (db) 5.18 8.75a, 7.95b 5.73 6.56 6.4 –
Volatiles matter 82.58 79.67a, 83.86b, 81.9c 73.74 75.99 82.5 42.2
Fixed carbon 8.97 8.68a, 10.78b, 12.6c 18.37 12.39 17.0 48.8
Ash 3.45 3.02a, 5.36b, 3.1c 2.21 5.33 0.5 5.8
Ultimate analysis (wt.% db)
C 46.62 48.79a, 49.07b, 53.78c 53.78 50.27 49.5 81.5
H 6.45 7.33a,6.48b,4.37c 7.20 7.07 6.2 4.0
N 1.21 0.00a,0.7b, 0.35c 0.00 0.42 0.4 1.2
S 0.035 0.68a,0.1b, 0.00c 0.51 0.63 – 3.0
O (by difference) 45.66 40.18a,38.29b, 41.5c 36.30 36.28 41.2 10.3
Calorific value (MJ/kg) 17.02 18.96a, 17.08b 22.14 20.64 19.2 25.0

Wb, wet basis; db, dry basis.


a
Ma and Yousof (2005).
b
Yang et al. (2006).
c
Abdullah and Gerhauser (2008).
d
Sukiran (2008).
e
Azali et al. (2005).
f
Demirbas (2004).
g
Bituminous coal.

TG (thermogravimetric)
DTG (differential thermogravimetric)
0.016
100
90 TG 0.014
DTG
80 0.012
Moisture
70 content Fixed
Weight loss, %

Volatile matters

dm/dT, mg/s
(Dry basis) content carbon 0.01
60
50 0.008

40 0.006
30
0.004
20
0.002
10
0 0
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

Fig. 1. Proximate analysis of EFB using thermogravimetry.

of fixed carbon and ash content at 8.79% and 3.45%, respectively, lower than that reported in the literature for different types of bio-
were present in EFB. This high content of volatile matter makes mass and coal (Ma and Yousof, 2005; Yang et al., 2006). Overall, the
EFB highly volatile and reactive (Demirbas, 2004) and very suitable ultimate analysis indicates that EFB is environmental friendly, with
for gas fuel production. The results from the proximate analysis are only trace amounts of nitrogen and sulfur. The H/C and O/C ratios
within the range of literature values. Generally, compared to other of EFB were 1.66 and 0.73, respectively. Considering only the main
types of palm oil biomasses such as shell and fiber, EFB has higher elements (C, H, and O), the molecular formula, based on C-atom, of
volatile matter and lower fixed carbon contents. The high ash value the EFB studied can be written as CH1.66O0.73.
of EFB compared to palm oil shell and woody biomasses is note- The EFB has an HHV of 17.02 MJ/kg, a value slightly lower than
worthy because ash is known to negatively affect the high heating that reported in the literature for wood and coal (Azali et al., 2005;
value (HHV) (Demirbas and Arin, 2002). Demirbas, 2004;Yang et al., 2006); however, the value is consistent
The ultimate analysis of EFB showed an oxygen, carbon, hydro- with the higher oxygen and lower carbon contents in the EFBs.
gen, nitrogen, and sulfur content of 45.66%, 46.62%, 6.50%, 1.00% Although the EFB heating value represents approximately 60% of
and 0.04%. These values are similar to those previously determined the heating value of coal, it is still a potential solid fuel to produce
(Abdullah and Gerhauser, 2008; Ma and Yousof, 2005; Sukiran, renewable energy.
2008; Yang et al., 2006), and in comparison to palm oil shell and
fiber, EFB has the highest oxygen and lowest carbon contents. This 3.1.2. Lignocellulosic composition
has a negative impact on the HHV because a higher oxygen concen- The results of the chemical analyses of EFBs are shown in Table 3.
tration (carbon–oxygen bonds) tends to decrease the calorific value The cellulose, hemicellulose and lignin contents of EFB were 22.24%,
of EFB (McKendry, 2002). The sulfur content in EFB was much 20.58%, and 30.45%, respectively. The cellulose value is much lower
632 M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636

Table 3
Lignocellulosic analysis of empty fruit bunch and solid waste biomasses.

Component EFB (this study) EFB (literature)a,b,c Oil palm shelld Oil palm fiberd Beech woode
Cellulose 22.24 38.1–63.0 20.8 33.9 45.3
Hemicellulose 20.58 20.1–35.3 22.7 26.1 31.2
Lignin 30.45 10.5–36.6 50.7 27.7 21.9
a
Law et al. (2007).
b
Kelly-Yong et al. (2007).
c
Abdulah and Gerhauser (2008).
d
Saka (2005).
e
Demirbas (2005).

Table 4
Physical and inorganic contents of empty fruit bunches, solid waste biomasses and coal.

Properties EFB (this study) EFB (literature) Oil palm shelld Oil palm fiberd Oak woode Coale,f
Fuel density (kg/m3) 1422 1420 1430 1400 750 1300
Inorganic content (% ash, mf)
Si2O 10.83 27.0a, 12.8b, 19.3c 2.96 3.54 49.0 59.7
K2O 53.73 44.0a, 49.1b, 34.7c 0.6 0.24 9.5 2.1
CaO 12.5 8.0a, 6.53b, 1.9c 0.48 1.15 17.5 2.1
MgO 8.75 4.8a 0.83 6.13 1.1 1.8
Cl 5.3 5.3a, 3.6b – – 0.8 –
Fe2O3 3.6 3.0a, 1.28b, 1.8c 0.08 0.24 8.5 8.3
Al2O3 1.22 0.97a, 0.46b, 1.2c 0.24 0.57 9.5 19.8
Na2O 1.54 0.55a, 1.25b, 0.8c – – 0.5 0.8
P2O5 1.84 3.6a, 1.12b 0.59 0.19 1.8 0.2

mf, Moisture free.


a
Lahijani and Zainal (2011).
b
Omar et al. (2011).
c
Yin et al. (2008).
d
Chaiyaomporn and Chavalparit (2010).
e
Demirbas (2004).
f
Bituminous coal.

than that found in the literature (Umikalsom et al., 1997; Kelly-Yong improves potassium devolatilization (Arvelakis et al., 2005). The
et al., 2007; Law et al., 2007), whereas the hemicellulose and lignin XRF analysis of the EFB ash showed 5.3% Cl, which is high com-
values are within the reported range (Table 3). EFB has a lower cel- pared to Cl content in wood. The tendency of a biomass to be sin-
lulose content than palm fiber and wood but a higher content than tered is indicated by the following parameters (Basu, 2006):
shells (Table 3). The hemicellulose value is lower than that found in
literature (Abdullah and Gerhauser, 2008; Kelly-Yong et al., 2007; K2 O þ Na2 O
>1 ð20Þ
Law et al., 2007), and the lignin content is much lower than that SiO2
found in shell and higher than that in fiber and wood (Table 3).
The high lignin content in EFB might result in a high bio-oil yield K2 O þ Na2 O
> 0:34 ð21Þ
using pyrolysis as proposed by Demirbas (2007) and a high tar con- HHV2
tent in the produced gas (Hu et al., 2006).

The biomass sintering tendency is severe when the fuel constit-


3.1.3. Physical and inorganic characteristics uents exceed this boundary condition. Because the sintering ten-
The physical and inorganic properties of EFB and their compar- dency value of EFB in this study exceeds the above boundary
ison with other types of palm oil wastes, wood and coal are pre- limits according to its measured inorganic contents, the agglomer-
sented in Table 4. Relative to wood and coal, the apparent ation problem was expected at high gasification temperatures;
densities of palm oil wastes are higher. The XRF analysis of the however, EFB contains high levels of MgO and CaO, which are
EFB (ash basis) showed that EFB has more potassium and sodium effective in reducing agglomeration (Lahijani and Zainal, 2011).
and less silica, calcium, iron, and aluminum than wood and coal. Certain pretreatment methods such as leaching and fractionation,
The high potassium content in EFB negatively affects the melting or use of bed additive materials, such as MgO, CaO, Fe2O3, dolo-
behavior of ash by decreasing its melting point, causing agglomer- mite, and limestone can effectively reduce the bed agglomeration
ation, formation of hard deposit at high temperatures potential (Basu, 2006).
(Obernberger and Thek, 2004). Moreover, alkali metals, such as Low levels of silica were found in EFB, with values comparable
potassium, react readily with silica sand, which is a common bed to those in the literature (Lahijani and Zainal, 2011; Omar et al.,
material and also used as the bed material in this gasification pro- 2011) and to those from other biomasses (Demirbas, 2004); how-
cess, by breaking the Si–O–Si bond and forming silicates. The ever, the values are higher than those from other palm oil wastes
silicates, which have melting points lower than 900 °C, tend to (Chaiyaomporn and Chavalparit, 2010). In the present study, the
form deposits on the reactor walls and create a sticky deposit on low silica content in the EFB, in association with the reasonable
the surface of the bed particles, causing bed agglomeration presence of some inorganic compounds, such as MgO, Fe2O3, and
(Arvelakis et al., 2005). Chlorine has also been reported as one of Al2O3, effectively alleviated the agglomeration tendency of EFB
the ash constituents that play a key role in the development of during gasification, but the presence of CaO (12.5%) might possibly
ash-related problems. This element along with potassium produces catalyze secondary reactions, causing an increase in hydrogen in
potassium chlorine which is stable at high temperatures and the produced gas, because CaO is a CO2 adsorbent.
M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636 633

TG (thermogravimetric)
DTG (differential thermogravimetric)
0.03
Stage 1 Stage 2 Stage 3
A
100 0.025
TG
DTG
80
Weight loss, %
0.02

dm/dT, mg/s
Lignin Decomposition
Lignin
Moisture and Char Combustion
60 Decomposition 0.015
Removal
Cellulose and
Hemicellulose
40 Decomposition 0.01

20 0.005

0 0
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

Fig. 2. Identification of decomposition stages of EFB by thermogravimetry.

(a) 10 C/min 20 C/min 30 C/min


100
Ts (initial reaction temperature)
90
Tmax1 (first peak temperature)
80
Tmax1 (second peak temperature)
Weight loss, %

70
Te (final reaction temperature)
60
50
40
30
20
10
0
0 100 200 300 400 500 600 700 800 900 1000

Ts Tmax 1 Tmax 2 Te

(b) 0.09
0.08
0.07
dm/dT, mg/min

0.06
0.05
0.04
0.03
0.02
0.01
0
0 100 200 300 400 500 600 700 800 900 1000
Tempertaure, °C

Fig. 3. Thermal degradation characteristics of EFB at different heating rates: (a) thermogravimetric analysis, (b) differential thermogravimetric analysis.

The char with the inorganic constituents formed during the 3.1.4. Thermal decomposition analysis
thermal treatment of EFB can be applied as soil conditioner be- As shown in Fig. 2, several stages of decomposition could be dis-
cause of its high nitrogen and potassium content (Omar et al., tinguished. At temperatures below 80 °C, no significant change or
2011). degradation was observed. In the first stage, a small DTG hump
634 M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636

(a) <0.3 mm 0.3-0.5 mm 0.5-1.0 mm

100
Ts (initial reaction temperature)
Tmax1 (first peak temperature)
80
Tmax1 (second peak temperature)

Weight loss, %
Te (final reaction temperature)
60

40

20

0
0 100 200 300 400 500 600 700 800 900 1000

Ts Tmax 1 Tmax 2 Te

(b) 0.03

0.025

0.02
dm/dT, mg/s

0.015

0.01

0.005

0
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

Fig. 4. Thermal degradation characteristics of EFB at different particle sizes: (a) thermogravimetric analysis, (b) differential thermogravimetric analysis.

Table 5 hemicellulose and cellulose (Lapuerta et al., 2004). The difference


Least-squares estimation results for different EFB particle sizes under air atmosphere. in the degradation processes of cellulose and hemicellulose is not
EFB particle size fully understood; nevertheless, hemicellulose has been established
break down at a lower temperature compared to cellulose
<0.3 mm 0.3–0.5 mm 0.5–1.0 mm
(Demirbas and Arin, 2002). Therefore, the loss of hemicellulose oc-
Moisture
curred first at a temperature range of 180–260 °C, followed by
Percentage (%) 2.5 2.0 1.5
degradation of cellulose at 260 °C to approximately 410 °C. Point
Cellulose and hemicellulose degradation
A in the DTG curve (Fig. 2) may represent the start of cellulose
Percentage (%) 47 50 52
A (S1) 4.0E+03 4.0E+03 6.5E+04 decomposition (Omar et al., 2011). More than 60% of the weight
E (kJ/mol) 61.17 61.56 73.76 loss of the sample occurred in this stage. The third stage of degra-
Lignin degradation and char combustion dation occurred between 410 and 680 °C and corresponds to lignin
Percentage (%) 43 40 38 decomposition and char oxidation (Lapuerta et al., 2004). The DTG
A (S1) 2.95 3.62 16.62 peak appearing around 460 °C is indicative of the decomposition of
E (kJ/mol) 40.06 41.51 47.99 lignin, whereas the small hump observed between 600 and 700 °C
Residue can be attributed to char combustion, which usually occurs
Percentage (%) 7.5 8.0 8.5 between 600 and 900 °C. The remaining material was considered
Standard deviation 0.078 0.065 0.098
ash.
A – Pre-exponential factor, E – activation energy. The extent of weight loss in the second and third stages of
decomposition depends on cellulose, hemicellulose, and lignin
contents in biomass (Lapuerta et al., 2004).
was observed around 100 °C, which is indicative of moisture re- The total weight loss of the EFB sample between 180 and 680 °C
moval. In the second stage between 180 and 400 °C, the chemical was more than 88%. The high content of volatile matter makes it
bonds in EFB began to break and the lightest volatile compounds easy to burn or gasify EFB compared to other types of biomasses,
were released. This stage usually represents the decomposition of such as oil palm shell, apricot stone, and coal (Yang et al., 2006).
M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636 635

3.2. Thermal conversion characteristic and kinetic behavior equilibrium with the temperature of the reactor or sample because
of heat transfer limitations (Wang et al., 2008). However, a higher
3.2.1. Effect of heating rate heating rate shortens the reaction time (in the same temperature
The TGA and DTG profiles at heating rates of 10, 20, and 30 °C/ region), and therefore, the temperature needed for the sample to
min in an air atmosphere are shown in Fig. 3. Both TGA and DTG decompose was higher. This caused the maximum rate curve to
curves shift to a higher temperature at higher heating rates. The shift to the right. This phenomenon has also been observed by
initial reaction temperature (Ts), the first DTG peak temperature others (Li et al., 2008; Orfao et al., 1999).
(Tmax1), the second DTG peak temperature (Tmax2) and the final
reaction temperature (Te) increased as the heating rates increased.
Lapuerta et al. (2004) and Li et al. (2008) have found similar results 3.2.2. Effect of particle size
for pine and corn straw pyrolysis, respectively. The Ts varied within The TGA and DTG profiles of different EFB particle sizes under
the temperature interval of 170–187 °C, Tmax1 is approximately air atmosphere and at a heating rate of 30 °C/min are compared
270–303 °C, Tmax2 was in the range of 423–445 °C and Te varied be- in Fig. 4. No significant change was observed for initial (Ts) and
tween 489 and 616 °C (Fig. 3). The lowest values (Ts = 170 °C, maximum temperature (Tmax1) at cellulose and hemicellulose
Tmax1 = 270 °C, Tmax2 = 423 °C, and Te = 489 °C) belong to the lower decomposition stage, the values of Ts and Tmax1 were approxi-
heating rate of 10 °C/min, whereas the upper values (Ts = 187 °C, mately 180 and 300 °C, respectively. As the size of the EFB particles
Tmax1 = 303 °C, Tmax2 = 445 °C, and Te = 616 °C) represent the high- increased, DTG peak temperature (Tmax2) and final reaction
est heating rate of 30 °C/min. With an increase in the heating rate, temperature (Te) decreased significantly. As shown in Fig. 4b, the
a large instantaneous thermal energy is provided to the system and peak temperatures Tmax2 were 477 °C for <0.3 mm, 474 °C for
a longer time may be required for the oxidation agent to reach 0.3–0.5 mm particles, and 464 °C for 0.5–1.0 mm particles.

EFB <0.3 mm
(a) 100
TG (Thermogravimetric) Model
0.6

80 TG (Thermogravimetric)-Experiment 0.5
% dry weight

DTG (Differential thermogravimetric)-Model 0.4


60

dm/dt
DTG (Differential Thermogravimertic)-Experiment
0.3
40
0.2
20 0.1
0 0
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

EFB 0.3-0.5 mm
(b) 100 TG (Thermogravimertic)-Model
0.8
0.7
80 TG (Thermogravimetric)-Experiment
% dry weight

0.6
DTG (Differential Thermogravimeric)-Model
60 0.5
dm/dt

DTG (Differential Thermogravimetric)-Experiment


0.4
40 0.3
0.2
20
0.1
0 0
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

EFB 0.5-1.0 mm
(c) 100 TG (Thermogravimetric)-Model 0.7
TG (Thermogravimetric)-Experiment 0.6
80
% dry weight

0.5
DTG (Differential Thermogravimetric)-Model
60 0.4
dm/dt

DTG (Differential Thermogravimetric)-Experiment


0.3
40 0.2
0.1
20
0
0 -0.1
0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

Fig. 5. Experimental and modeled TG and DTG curves, different EFB particle sizes under air atmosphere.
636 M.A.A. Mohammed et al. / Bioresource Technology 110 (2012) 628–636

Similarly, the final reaction temperature Te was 642, 591, and Chaiyaomporn, K., Chavalparit, O., 2010. Fuel pellets production from biodiesel
waste. EnvironmentAsia 3, 103–110.
566 °C for the particles size of <0.3, 0.3–0.5 and 0.5–1.0 mm,
Demirbas, A., 2004. Combustion characteristics of different biomass fuels. Progress
respectively. in Energy and Combustion Science 30, 219–230.
Demirbas, A., 2005. Pyrolysis of ground beech wood in irregular heating rate
3.3. Kinetic analysis of EFB gasification using the least-squares conditions. Journal of Analytical and Applied Pyrolysis 73, 39–43.
Demirbas, A., 2007. The influence of temperature on the yields of compounds
estimation model existing in bio-oils obtained from biomass samples via pyrolysis. Fuel
Processing Technology 88, 591–597.
Table 5 represents the calculated activation energy E and pre- Demirbas, A., Arin, G., 2002. An overview of biomass pyrolysis. Energy Sources 24,
471–482.
exponential factors A of the least-squares estimation model, in Harris, A.T., Zhong, Z., 2008. Non-isothermal thermogravimetric analysis of plywood
addition the standard deviation between the experimental and wastes under N2, CO2 and O2 atmospheres. Asia-Pacific Journal of Chemical
modeled curve is also included. The values of A change for each E Engineering 3, 473–480.
Hu, G., Xu, S., Li, S., Xiao, C., Liu, S., 2006. Steam gasification of apricot stones with
value. The change in activation energies during the heating process olivine and dolomite as downstream catalysts. Fuel Processing Technology 87,
indicates that the degradation process of EFB was much more com- 375–382.
plex than that of a single component. The kinetic analysis of holo- Idris, S.S., Rahman, N.A., Ismail, K., Alias, A.B., Rashid, Z.A., Aris, M.J., 2010.
Investigation on thermochemical behaviour of low rank Malaysian coal, oil
cellulose (cellulose and hemicellulose) decomposition stage palm biomass and their blends during pyrolysis via thermogravimetric analysis
showed that the activation energies were 61.17, 61.56, and (TGA). Bioresource Technology 101, 4584–4592.
73.22 kJ/mol for particle sizes <0.3, 0.3–0.5, 0.5–1.0 mm, respec- Kataki, R., Konwer, D., 2001. Fuelwood characteristics of some indigenous woody
species of north-east India. Biomass and Bioenergy 20, 17–23.
tively, whereas for the combustion stage of lignin and char, the
Kelly-Yong, T.L., Lee, K.T., Mohamed, A.R., Bhatia, S., 2007. Potential of hydrogen
activation energies were 40.06, 41.5, and 50.20 kJ/mol for the same from oil palm biomass as a source of renewable energy worldwide. Energy
particle size category. The activation energies and the pre- Policy 35, 5692–5701.
exponential factors increased with increasing EFB particle size for Lahijani, P., Zainal, Z.A., 2011. Gasification of palm empty fruit bunch in a bubbling
fluidized bed: a performance and agglomeration study. Bioresource Technology
both pyrolysis and gasification stages (Table 5). The experimental 102, 2068–2076.
and modeled TGA and DTG curves for different particle sizes of Lapuerta, M., Hernández, J.J., Rodríguez, J., 2004. Kinetics of devolatilisation of
EFB are presented in Fig. 5. Both the experimental and modeled forestry wastes from thermogravimetric analysis. Biomass and Bioenergy 27,
385–391.
curves for the EFB sample heated at three different heating rates Law, K.N., Daud, W.R.W., Ghazali, A., 2007. Morphological and chemical nature of
showed a very good fit of the model to the experimental results. fiber strands of oil palm empty-fruit-bunch (OPEFB). BioResources 2,
The standard deviation of less than 0.1% validates the good agree- 351–362.
Li, Z., Zhao, W., Meng, B., Liu, C., Zhu, Q., Zhao, G., 2008. Kinetic study of corn straw
ments between the experimental and the modeled values. pyrolysis: comparison of two different three-pseudocomponent models.
Bioresource Technology 99, 7616–7622.
Ma, A.N., Yousof, B., 2005. Biomass energy from palm oil industry in Malaysia.
4. Conclusions
Ingenieur 27, 18–25.
McKendry, P., 2002. Energy production from biomass (Part 2): Conversion
EFB has a lower heating value than wood and coal due to the technologies. Bioresource Technology 83, 47–54.
higher contents of moisture and oxygen, but the high volatile mat- Mohammed, M.A.A., Salmiaton, A., Wan Azlina, W.A.K.G., Mohamad Amran, M.S.,
fakhr’l-Razi, A., Tawfiq-Yap, Y.H., 2011. Hydrogen rich gas from oil palm
ter content compensates for this disadvantage by boosting the biomass as a potential source of renewable energy in Malaysia. Renewable and
reactivity of the EFB. Nevertheless, high potassium content causes Sustainable Energy Reviews 15, 1258–1270.
ash-related problems in the gasification system at high tempera- Obernberger, I., Thek, G., 2004. Physical characterisation and chemical composition
of densified biomass fuels with regard to their combustion behaviour. Biomass
tures. Three decomposition stages were observed during the ther- and Bioenergy 27, 653–669.
mal degradation of EFB. Assuming three independent parallel Omar, R., Idris, A., Yunus, R., Khalid, K., Aida Isma, M.I., 2011. Characterization of
reactions, the least-squares estimation method for a first-order empty fruit bunch for microwave-assisted pyrolysis. Fuel 90, 1536–1544.
Orfao, J.J.M., Antunes, F.J.A., Figueiredo, J.L., 1999. Pyrolysis kinetics of
reaction model was successfully applied to describe the EFB gasifi- lignocellulosic materials – three independent reactions model. Fuel 78, 349–
cation, which showed a good fit with experimental data. 358.
Quan, C., Li, A., Gao, N., 2009. Thermogravimetric analysis and kinetic study on large
particles of printed circuit board wastes. Waste Management 29, 2353–2360.
Acknowledgements
Saka, S., 2005. Whole efficient utilization of oil palm to value-added products. In:
Proceedings of JSPS-VCC Natural Resources & Energy Environment Seminar.
The author’s would like to thank Department of Chemical & Shen, D.K., Gu, S., Luo, K.H., Bridgwater, A.V., Fang, M.X., 2009. Kinetic study on
thermal decomposition of wood in oxidative environment. Fuel 88, 1024–1030.
Environmental Engineering, Faculty of Engineering, Universiti Pu-
Sonobe, T., Worasuwannarak, N., 2008. Kinetic analyses of biomass pyrolysis using
tra Malaysia for financial support on this research. the distributed activation energy model. Fuel 87, 414–421.
Sukiran, M., 2008. Pyrolysis of Empty Oil Palm Fruit Bunch using the Quartz
References Fluidized-fixed Bed Reactor. Master Thesis, University of Malaya.
Umikalsom, M.S., Ariff, A.B., Zulkifli, H.S., Tong, C.C., Hassan, M.A., Karim, M.I.A.,
1997. The treatment of oil palm empty fruit bunch fibre for subsequent use as
Abdullah, N., Gerhauser, H., 2008. Bio-oil derived from empty fruit bunches. Fuel 87, substrate for cellulase production by Chaetomium globosum Kunze. Bioresource
2606–2613. Technology 62, 1–9.
Arvelakis, S., Gehrmann, H., Beckmann, M., Koukios, E.G., 2005. Preliminary results Wang, G., Li, W., Li, B.Q., Chen, H.K., 2008. TG study on the pyrolysis of biomass and
on the ash behavior of peach stones during fluidized bed gasification: its three components under syngas. Fuel 87, 552–558.
evaluation of fractionation and leaching as pre-treatments. Biomass and Yang, H., Yan, R., Chen, H., Lee, D., Liang, D., Zheng, C., 2006. Pyrolysis of palm oil
Bioenergy 28, 331–338. wastes for enhanced production of hydrogen rich gases. Fuel Processing
Azali, A., Nasrin, A.B., Choo, Y.M., Adam, N.M., Sapuan, S.M., 2005. Development of Technology 87, 935–942.
gasification system fuelled with oil palm fibres and shells. American Journal of Yin, C.Y., Kadir, S.A.S.A., Lim, Y.P., Syed-Ariffin, S.N., Zamzuri, Z., 2008. An
Applied Sciences, 72–75 (special issue). investigation into physicochemical characteristics of ash produced from
Basiron, Y., 2007. Palm oil production through sustainable plantations. European combustion of oil palm biomass waste in a boiler. Fuel Processing Technology
Journal of Lipid Science and Technology 109, 289–295. 89, 693–696.
Basu, P., 2006. Combustion and Gasification in Fluidized Beds. CRC Press, Boca
Raton.

You might also like