Você está na página 1de 8

J. Anal. Appl.

Pyrolysis 85 (2009) 155162

Contents lists available at ScienceDirect

Journal of Analytical and Applied Pyrolysis


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

Thermal decomposition study on Jatropha curcas L. waste using TGA and xed
bed reactor
Viboon Sricharoenchaikul a, Duangduen Atong b,*
a
b

Department of Environmental Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
National Metal and Materials Technology Center, 114 Thailand Science Park, Pathumthani 12120, Thailand

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 27 May 2008
Accepted 25 November 2008
Available online 3 December 2008

Pyrolysis experiments on Jatropha curcas L. (physic nut) waste were carried out using thermogravimetric
analysis (TGA) and a xed bed quartz reactor to determine suitable degradation model as well as
investigate the effect of operating conditions on product distribution. It was found that the main thermal
decomposition of physic nut waste generally occurred over the temperature range of 250450 8C. The
three-parallel reactions model was applied for simulating the degradation of this waste. The model
agreed relatively well with the experimental data. From the model, the activation energy of
hemicelluloses, cellulose and lignin was in the range of 4168, 187235, and 97150 kJ/mol,
respectively. Reaction orders of those fractions were in the range of 2.43.2. Results from pyrolysis
process using xed bed reactor indicated that increase in temperature and hold time lead to greater
production of hydrogen, methane and light hydrocarbons with highest gas production detected at
900 8C. Tar decomposed at higher temperatures resulted in lower liquid yield while gas yield and total
conversion increased. Liquid product consists of several fatty acids such as palmitic acid, stearic acid, and
oleic acid in the range of 1023%, 512%, and 3542%, respectively. The amount of char residue decreased
with increasing reactor temperature and hold time. Fixed carbon in char increased with temperature
with the expense of volatile matter while there was little change on ash content. Generally, pyrolysis of
this residue may be applied for the production of value-added products as well as fuels after some
upgrading processes.
2008 Elsevier B.V. All rights reserved.

Keywords:
Pyrolysis
Physic nut
Thermogravimetric analysis
Model

1. Introduction
One of the most viable renewable energy sources is biomass
from agricultural residues. It is the third largest primary energy
source after coal and oil [1]. The potential of biomass to help to
meet the world energy demand has been widely recognized. At the
time of soaring prices amid limited supply of fossil fuels, the use of
biomass as an alternate energy source is especially important for
most developing countries where their economies are based
largely on agricultural and forestry sectors. Biochemical and
thermochemical processes may be used for the recovery of valueadded products and energy from biomass. Thermochemical
conversion of biomass such as pyrolysis yields liquid oil, solid
charcoal, and gases in different proportions, depending on the
temperature, heating rate, reactor type, and biomass particle size.
Recent researches focus on the distribution of various gas products
from pyrolysis of biomass, aiming to improve the yield of clean

* Corresponding author. Tel.: +66 2 564 6500x4230; fax: +66 2 564 6501.
E-mail address: duangdua@mtec.or.th (D. Atong).
0165-2370/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jaap.2008.11.030

energy such as hydrogen for utilization in hydrogen-powered


engine as well as fuel cell [26].
Physic nut or Jatropha Curcas L. is a plant oil similar to palm
oil. Jatropha plant has potential as a renewable energy crop as its
oil may be used directly with slow speed diesel engine or
upgraded via transesterication to conventional biodiesel.
Extraction of physic nut oil results in residue that is needed to
be disposed. Generally, collection and disposal of residues are
becoming more difcult and expensive and may create
environmental problems if not properly done. Pyrolysis could
be a promising residue management option to convert this waste
to fuel products that are easier to transport, storage, handling,
and utilizing. However, the information on detailed kinetic of
pyrolysis operation for this waste is not readily available which
would prevent the effective design, model, and optimization of
this promising waste utilization route. As a result, the aim of this
work is to study the thermal degradation kinetics of physic nut
waste and simulate its decomposition prole using appropriate
model. In addition, the characteristic of products obtained from
fast pyrolysis of this residue in a xed bed quartz reactor is also
examined.

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

156
Table 1
Chemical analysis of physic nut waste.

wt.%
Elemental analysis
Carbon
Hydrogen
Nitrogen
Potassium
Calcium
Phosphorus
Magnesium
Sulfur
Chlorine
Oxygena

52.30
6.50
5.20
3.84
2.00
1.98
1.21
0.58
0.21
26.18

Proximate analysis
Volatiles
Fixed carbon
Ash
Moisture

79.20
18.86
1.50
0.44

Component analysis
Hemicellulose
Cellulose
Lignin
Othera

17.47
56.31
23.91
2.31

By difference.

separate slopes of constant mass degradation. TGA was also used


for proximate analysis including moisture, volatiles, xed carbon,
and ash content. The apparatus consists of a microbalance within a
furnace, allowing the weight of the sample to be continuously
monitored while the temperature is controlled. Samples can be run
either in a dynamic (temperature ramp) or in an isothermal mode.
Once these data are collected, analysis of the raw numbers leads to
the kinetic rate parameters: n, order of reaction; Ea, activation
energy; and A, pre-exponential factor [715].
In this work, experiments were carried out in inert gas (N2,
99.9995%) with ow rate of 50 ml/min. Sample taken was 40
45 mg for each run. Dynamic trials were performed at heating rate
of 5, 10, 15, 30 and 90 8C/min from ambient to 900 8C. Prior to each
experiment, mass and temperature modules of the analyzer were
calibrated to obtain reliable and reproducible mass and temperature data. Mass losses that correspond to temperature change were
continuously recorded with data acquisition working in coordination with the furnace and control unit of the analyzer. Variation of
sample mass with respect to temperature change (TG data) and its
rst derivative (DTG data) was continuously collected.
The derived TG data was then tted with a few selected models
to test the predictive capability of each model by using
optimization function in MATLAB1 program. The kinetic parameters were then followed from the model tting results.

2. Experimental procedure

2.3. Model description

2.1. Raw material

In this work, the model for kinetics of physic nut waste


pyrolysis was chosen to be of parallel reaction type by examination
of effects of heating rate on the shape of obtained thermograms.
Font et al. [15] applied two-parallel reactions model in their work
for describing the pyrolysis scheme of almond shell. This was
accomplished by assuming that the raw material consists of two
independent fractions which disintegrate at different rates and
temperatures, creating two main competing reactions which occur
simultaneously. The rst-order kinetics for both competing
reactions was assumed in their study. In order to increase
exibility of the proposed three-parallel reactions model, the
order of reaction was set as a free parameter though identical for all
components for the work reported here.
The reaction starts with the conversion of raw material which
consists of three homogeneous matters and each component
decomposes simultaneously at different rate and temperature,
producing volatile matters and solid char. The overall reaction is
expressed in Eq. (1) along with each individual parallel reaction. It
is assumed that each reaction is kinetically controlled and there are
no secondary reactions among the released volatile products.

Physic nut waste sample was acquired from local oil extraction
plant. It was rst separated from physical impurities and oven
dried overnight at 80 8C. Samples were then sieved to size fraction
of 0.430.50 mm and kept in desiccators. Results of elemental and
proximate analyses of samples are presented in Table 1. Elemental
analysis reports the C, H, N, O and other elements in the sample.
Proximate analysis classies the sample in terms of moisture,
volatile matter, xed carbon and ash. The volatile matter mainly
consists of organic compounds. Moisture content determined by
the proximate method represents the water that is physically
bound to the matrix structure; water released by chemical
reactions during pyrolysis is classied with the volatiles. The
ash content is determined by combustion of the volatile and xed
carbon fractions. Additional chemical components of physic nut
waste sample were also identied using Tappi T203 and T222
standard methods and shown in the same table, which are
hemicellulose, cellulose and lignin.
For physic nut waste, carbon is the main element with
52.30 wt.%. Large amount of oxygen is reported by difference, so
this number is not conclusive since there may be other elements
not analyzed by this method. The volatile matter of 79.20 wt.%
mainly consists of various organic compounds. Low ash content is
an indicator that physic nut waste should be a good candidate for
the production of fuel via thermochemical process. Generally, data
obtained are of typical range for biomass samples.
2.2. Characterization of thermal degradation process
The thermal degradation characteristic of physic nut waste was
studied using thermogravimetric analyzer (Mettler Toledo: TGA/
SDTA 851e). Thermogravimetric analysis (TGA) is one of the most
common techniques for the investigation of thermal events and
their associated kinetics during pyrolysis of solid raw materials
such as coal, biomass, and plastic. It provides a measurement of
weight loss of the sample as a function of time and temperature.
The kinetics of these thermal events has been determined by the
application of the Arrhenius equation corresponding to the

Raw material ! volatile1 volatile2 volatile3 char


k1

M1 ! volatile1 char
k2

(1)

M2 ! volatile2 char
k3

M3 ! volatile3 char
where M1, M2, and M3 are each component in raw material which
independently decompose at different rate constant of k1, k2, and
k3, respectively.
Assume that the rate of decomposition can be expressed by one
kinetic scheme as
da
kT f a
dt

(2)

where


E
kT A exp 
RT

(3)

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

157

or


da
E
f a
A exp 
RT
dt

(4)

Here, k(T) is the kinetic rate constant, n is the order of the reaction,
A is a pre-exponential factor, E is the activation energy, R is the gas
constant, t is time, and a is the normalized fractional conversion
and is dened as

w0  w
w0  w1

(5)

where w0 , w and w1 are the initial, instantaneous, and nal mass of


the sample, respectively.
Eq. (2) may be written as
da
kT

dT
f a
b

(6)

Replace k(T) with Eq. (3) then rearrange and integrate to yield


Z
da
A T
E
dT
(7)

exp 
RT
b 0
a;b;c f a

Z a

where a, b and c are initial values of a1, a2, and a3, respectively.
They indicate the initial weight fractions of components 1, 2, and 3
in the starting raw material and are assumed to be constant and
depend only on the characteristics of the raw material. For physic
nut waste, these are cellulose, hemicelluloses and lignin. The
kinetic parameters are assumed to be the linear summation of
kinetic value of each component. The relationship between a, b,
and c is expressed as
abc 1

(8)

For n order reactions, f(a) may be described as


f a 1  an

(9)

Substitution of Eq. (9) into Eq. (7) then integrate and combine those
three expressions to yield
"

#1=1n




n  1A1 RT 2
2RT
E1
1n
a
a
1
exp 
E1
RT
bE1
"
#1=1n




n  1A2 RT 2
2RT
E2
1n
b

1
exp 
E2
RT
bE2
"
#1=1n



2
n  1A3 RT
2RT
E3
1n
c

1
exp 
E3
RT
bE3
(10)
The experimental data is then used to t with Eq. (10). Kinetic
parameters, A, E, and n, are determined by minimizing the sum of
square of relative error (SSRE), dened in Eq. (11), through the use
of MATLAB1 programming.
SSRE

X aexp  acal 2

aexp

(11)

Here, aexp and acal are the experimental and simulated fractions of
char, respectively. The accuracy of the model also depends on the
size of time step and numerical techniques used in the calculation
as well as some input parameters such as initial mass fraction of
each components in physic nut waste which must be experimentally determined.
2.4. Fast pyrolysis experiment
Pyrolysis experiments were conducted in a quartz tube reactor
(25 mm id. and 300 mm long). Reactor and other apparatus used

Fig. 1. Schematic of xed bed pyrolysis system.

in this setup were displayed in Fig. 1. During any experiments, the


reactor was preheated to the nal set point using an automatic PID
temperature controller. Purging of the system with argon was
then carried out with electronic mass ow controller to ensure
inert condition inside the reaction chamber. The reaction started
as soon as about one gram of physic nut waste was placed into a
specially fabricated fritted disc quartz sample retainer and rapidly
inserted into the center of the tubular reactor. Temperature
measurement was taken with a K-type thermocouple placing in
the middle of tube reactor to keep isothermal condition
throughout each trial. The feed rate of argon was 67.5 ml/min
while the output gas ow rate was registered using a rotameter.
Three stage condensers are utilized for trapping of condensable
products.
Experimental conditions were designed as follows: nal
pyrolysis temperature of 500, 700, and 900 8C, respectively, each
with hold times of 3, 9, and 15 min. After each run, char left in
sample retainer was carefully weighed and kept for further
analysis. Liquid product was collected in cold traps maintained at
about 0 8C. Dichloromethane was used to wash any liquid
condensed in gas transfer line and mixed with liquid from
condensers for analysis. Non-condensable gas was collected in
Tedlar1 sampling bags. This gas product was analyzed using a gas
chromatograph (Varian, CP-3800) with a thermal conductive
detector (TCD) and ame ionized detector (FID). The two columns
are: (1) molecular sieve 5A (MS-5A), for the analysis of H2 and CH4
and (2) Porapak Q (PPQ), for the identication of other light
hydrocarbons.
Liquid product was analyzed using a gas chromatograph
(Agilent Technology, 6890N) with a FID and DB-WAX capillary
column. Only fatty acid components were investigated by
conversion to their respective methyl esters. Fatty acid methyl
esters were prepared by transesterication and puried by thinlayer chromatography using potassium hydroxide for the saponication of glycerides in a reux condenser attached to the ask and
heated to about 200 8C until the fat droplets disappeared. Methanol
and heptane were added through the condenser and the boiling
continued for some minutes. After addition of sodium chloride, the
obtained heptane solution was prepared for direct injection into
gas chromatography.
Solid and liquid yields were calculated on a dry basis from the
measured weight of each fraction, while the gas yield was
evaluated by difference. Yields of identied gas were calculated

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

158

Fig. 4. Mass loss fraction vs time for experimental and simulated data at various
heating rates.
Fig. 2. Mass loss prole of physic nut waste at various heating rates.

based on percentage of mass input as mass of that particular gas


species.
3. Results and discussion
Thermochemical conversion of physic nut waste was performed at various heating rates to study kinetic parameters of the
pyrolysis process. The obtained data were then tted with threeparallel reactions model proposed in this study. Additional results
on product yields of fast pyrolysis process of the same raw material
were also presented in the following discussion.
3.1. Thermogravimetric study
The thermogravimetric data (TG) and its rst derivative data
(DTG) of physic nut waste at different heating rate are plotted
against furnace temperature at various heating rate in Figs. 2 and 3,
respectively. Generally, nal char yields at 900 8C varied with
heating rates from 15.2% to 23.4% at 590 8C/min, respectively.
Slower heating rates yield better separation among decomposition
stages of physic nut waste components as seen in DTG curve.
In order to understand the degradation behavior of biomass
such as physic nut waste, separate components of this residue
should be examined. Actual chemistry of biomass can be very
complicated. Simplication may be done by considering only
major constituents such as hemicellulose, cellulose and lignin
according to their mass loss curve. Thermal degradation of these
individual components may be superimposed to simulate the
overall degradation of the original biomass. It is known from
previous studies that thermal decomposition of hemicellulose and

cellulose have the rst (150350 8C) and second weight loss (275
350 8C) steps for lignocellulosic materials after the initial weight
loss (30150 8C) associated with the moisture loss. However, lignin
undergoes gradual decomposition over a wide temperature
interval (3001000 8C) [1620]. Decomposition of raw material
during these initial steps results in the evolution of light volatile
compounds earlier detected during each run.
Maximum weight losses occur from 250 to 450 8C (Fig. 2) for
physic nut waste. This indicates that there is major reaction
scheme occurring during the pyrolysis process for this type of
material. The maximum peak of DTG curve is probably contributed
by the decomposition of the lighter fraction and the small shoulder
corresponds to the decomposition of the heavier component. At
lower heating rate, it is obvious that the major degradation of
physic nut waste composes of three distinct peaks with partial
overlapping. This result supports the formulation of three-parallel
reactions model applied in this work. These plots exhibit right shift
of the volatile evolution curves with increase of heating rate.
Fusion of devolatilization events is also observed at higher heating
rate of 30 and 90 8C/min due to competing effects of heat and mass
transfer to the material. At high heating rate, it is possible that heat
cannot effectively transfer into the inside of raw material resulting
in greater temperature gradient between inner and outer layer of
waste particles. Since the particle used in this experiment is not
very ne, the lag between measured surrounding temperature and
that of inside particle may occur. Generally, higher reaction rates
(greater slope) were also noticed at higher heating rate.
Data on thermal decomposition of physic nut waste at four
different heating rates were used to investigate kinetic parameters
of this pyrolysis process. The experimental data of physic nut was
tted with the three-parallel reactions model described earlier.
Comparisons between experimental and simulated data using this
model are shown in Figs. 4 and 5 for various heating rates. The
results show good agreement between the predicted results of
Table 2
Kinetic parameters of three-parallel reactions model for pyrolysis of physic nut
waste.
Kinetic parameters

Fig. 3. DTG plot of physic nut waste at various heating rates.

a
b
c
A1 (min1)
E1 (kJ/mol)
A2 (min1)
E2 (kJ/mol)
A3 (min1)
E3 (kJ/mol)
n

Heating rate (8C/min)


5

15

30

5.93  103
68.0
8.35  1014
235.3
5.81  1016
149.5
3.24

0.342
0.450
0.208
3
4.32  10
1.76  103
49.3
45.7
2.56  1014
2.78  1015
227.4
218.5
4.92  1014
2.21  1014
143.2
98.7
3.05
2.37

90

2.06  102
40.9
1.92  1015
186.5
1.56  1015
97.3
2.46

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

159

Fig. 5. Mass loss fraction vs temperature for experimental and simulated data at heating rate of (a) 5 8C/min, (b) 15 8C/min, (c) 30 8C/min, and (d) 90 8C/min.

three-parallel reactions model and acquired data. The model


agrees well with the experimental data within 4% of deviation for
all heating rate condition being studied here.
The three-parallel reactions model consists of several kinetic
parameters, some parameters are determined experimentally,
such as the initial fraction of each starting components. Fitted
kinetic parameters, order of reaction (n), activation energy (E) and
frequency factor (A), determined from the model are listed in
Table 2. Similar to typical bio-based materials, the tted activation
energy of rst fraction (E1) is for degradation of hemicelluloses
(4168 kJ/mol), the activation energy of second fraction (E2) is
likely for that of celluloses (187235 kJ/mol) and the last fraction
(E3) would be that of lignin and others (97150 kJ/mol). The order
of reaction is in range of 2.43.2. The activation energy of physic
nut waste appears to decrease with increasing of heating rate while
the trend of frequency factor is not clear and varies with different
components. The frequency factor and activation energy indicate
how fast and easy for the pyrolysis reaction may proceed. The
higher frequency factor and lower activation energy, the faster and
easier would be for the pyrolysis reaction to occur. The order of the

reaction decreases with increasing of heating rate for physic nut


waste.
Due to the dissociation of relatively weak bond between carbon
and oxygen atoms to pentosan in and stronger hydrogen bonding,
the activation energy of hemicellulose is lower than that of
cellulose. Comparison on the calculated kinetic data from physic
nut waste to those obtained from other biomass with relatively
similar experiments is summarized in Table 3 as reported by
Caballero et al. [16], Manya` et al. [20], Hu et al. [21], and Miranda
et al. [22]. There is obviously large diversity among these kinetic
values due to difference in lignocellulosic species of each biomass.
In this context, cotton seems to be more similar to physic nut waste
than any other biomass. Generally, the acquired data are within the
range typical for biomass.
3.2. Product yields
In order to achieve a more practical utilization of physic nut
waste, experiments were carried out to study effect of operating
condition on product yields from fast pyrolysis of this residue.

Table 3
Kinetic parameters for pyrolysis of different biomass.
Sample

Heating rate (8C/min)

Temperature (8C)

Caballero et al. [16]

Almond shell

10

100700

Manya` et al. [20]

Bagasse

Hu et al. [21]

Rice husk

Miranda et al. [22]

This work

Components
1st fraction

2nd fraction

3rd fraction

A1 = 5.9  1013
E1 = 174.1

A2 = 7.5  1013
E2 = 192.1

A3 = 3.0  1013
E3 = 193.5

251000

A1 = 4.7  1015
E1 = 194.0

A2 = 1.0  1018
E2 = 243.3

A3 = 0.7
E3 = 37.0

30900

A1 = 9.6  1012
E1 = 162.0

A2 = 2.6  1015
E2 = 200.0

A3 = 8.0  101
E3 = 32.0

Cotton

251000

A1 = 1.5  103
E1 = 40.0

A2 = 4  1015
E2 = 190.0

A3 = 9  1013
E3 = 186.0

Physic nut waste

25900

A1 = 5.93  103
E1 = 68.0

A2 = 8.35  1014
E2 = 235.3

A3 = 5.81  1016
E3 = 149.5

160

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162


Table 4
Analysis of char from pyrolysis of physic nut waste at 15 min hold time.
Temperature (8C)
500

700

900

Elemental analysis (wt.%)


C
H
N
S
Oa

36.55
0.83
1.30
0.43
59.67

36.93
0.58
1.16
0.41
57.45

39.41
0.25
0.59
0.35
55.48

Proximate analysis (wt.%)


Volatile
Fixed carbon
Ash

50.93
11.27
1.22

23.12
34.33
3.47

12.06
40.31
3.92

Fig. 6. Product yields from fast pyrolysis of several agricultural residues.

Products from biomass pyrolysis are inuenced by several


parameters, such as temperature, residence time, heating rate,
reactor type as well as biomass size and species. Since temperature
and residence time play major role on biomass pyrolysis, many
researchers focused on these effects in their studies using both
uidized bed and xed bed reactors [2328]. In this work, physic
nut residue rapidly decomposed to gas, liquid, and solid fractions
as soon as being inserted into reaction chamber. Fig. 6 displays the
overall distribution of these fractions obtained from trials at the
temperature of 500900 8C along with data from other researchers
using fast pyrolysis process to convert some agricultural residues.
The results indicate that raising temperature leads to an increase in
the production of gas and decrease in more condensed fractions.
The decrease in liquid and char yields with temperature is mainly
attributed to an increase in the devolatilization of the organic
material, although partial pyrolysis of the char may also occur to
some extent.

By difference.

It may be noticed from the comparison of product yields from


physic nut waste with those of rice husk [17], olive husk [29] and
cotton cocoon shell [30] that solid yields from physic nut waste are
much higher than other residues. Liquid yields are relatively
similar for all residues with lower gas yields for physic nut waste,
especially at higher temperature. This is partly due to difference in
composition of raw materials. In addition, higher solid yield from
physic nut waste is probably due to the heat transfer limitation of
the xed bed pyrolysis setup of this work which could not heat the
sample fast enough so ash pyrolysis could take place instantly.
Caglar and Demirbas [29,30] used stainless wired mesh to feed
smaller sample size of 0.060.15 mm and lesser sample amount
into heated furnace so the onset of pyrolysis could occur rapidly
resulting in higher liquid and gas yields. This is the same for the
work by Williams and Nugranad [17] who used uidized bed
system known for superior heat transfer characteristic, so the
inception of pyrolysis for their case could also rapidly proceed.
Fixed bed reactor which may not heat sample as rapid as those
works was used. Lager sample size (0.50 mm) and sample amount
when compared to reactor size were used in each run which
probably cause uneven heating throughout sample mass. As a
result, the onset of pyrolysis may not be as rapid and lead to more
solid yield.
The chemical analysis of char obtained from pyrolysis of physic
nut waste at different temperatures for 15 min hold time is shown
in Table 4. As expected, an increase in temperature results in char
with higher xed carbon and decrease in volatile matter. Elemental
analysis of char shows similar trend of higher portion of carbon
content and lower amount of other elements with temperature.
Gas products were characterized by gas chromatography
equipped with FID and TCD. FID is sensitive for characterizing
hydrocarbon compounds while TCD is used for the analysis of
hydrogen. Quantitative determination of hydrocarbon gases range
from C1 to C4 was accomplished by comparing the GC response
against standard mixed gases.
Analysis of H2 in product gas as amount of hydrogen in raw
material that is converted to hydrogen gas is displayed in Fig. 7 for
pyrolysis at various time and temperature. Highest conversion to
H2 was achieved at higher temperature of 900 8C where 3.25%
conversion was registered. This amount is relatively small which is
typical for pyrolysis process where availability of free hydrogen
radical precursor is scarce. From instantaneous hydrogen production prole, the highest hydrogen conversion occurs around
4.5 min of reaction and almost complete after 7.5 min. For practical
purpose, this means that the residence time of 57 min for
pyrolysis reaction of this waste using xed bed reactor would be
ample if the purpose is for hydrogen production. Though, this is
usually not an objective for pyrolysis operation.
Other detected hydrocarbons in gas product including methane
and other C1C4 are analyzed and reported in Figs. 8 and 9 as

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

Fig. 7. Hydrogen gas produced from pyrolysis of physic nut waste.

Fig. 8. Methane produced from pyrolysis of physic nut waste.

amount of carbon in raw material converted to each specied gas


species. At higher temperatures, methane quickly increased to
maximum values and then rapidly decreased with hold time as raw
material depleted. Final yield of methane obtained from the
accumulation plot at the longest hold time indicated that its
production more than doubled from 2.1% to 4.7% as temperature
increased from 500 to 900 8C. The range of conversion to methane
from pyrolysis of physic nut waste in this work is quite common to
those of other agricultural residues.
Due to complex nature of hydrocarbon gas from pyrolysis of any
biomass, typical GC results reveal more than 20 peaks of gas

161

Fig. 10. Relative distribution of fatty acids in liquid product.

species for each sample. The analysis of these gases was performed
to acquire only the range of carbon number in the sample without
any effort to individually identify and quantify each gas. Peak area
of these gases which response to each carbon number was
approximately quantied against the standard of methane,
ethylene, propane, and butane for C1, C2, C3, and C4, respectively.
Analyses of these hydrocarbons were then collectively plotted as
C1C4 species (Fig. 9). Overall change is quite similar to that of
hydrogen and methane where greater formation occurs at higher
temperature though the conversion to these gases is much higher
than hydrogen and methane. At 500 8C and long hold time, the
conversion of carbon in raw material to C1C4 species is 28.5% and
rises to 48.9% at 900 8C. It should be noted that the distribution
among each C1C4 species is greatly different (not shown here).
Most of the detected hydrocarbon gases are those of C4 species
which amount to more than 20% on any trials and maximum at
45.6% for pyrolysis at 900 8C and 13.5 min hold time. In any
experiments, the release of C2 species is always lower than 1.0%
while those of C3 species are inconsistent. Some C5 species are also
detected but not reported here due to their very low quantities.
Liquid products were further processed and analyzed for fatty
acid methyl esters per procedure described earlier. The relative
distribution of fatty acids was determined as percent of
chromatographic peak area of the following derivative acids:
palmitic acid (C16:0:C16H32O2), palmitoleic acid (C16:1:C18H34O2),
stearic acid (C18:0:C18H36O2), oleic acid (C18:1:C18H34O2), linoleic
acid (C18:2:C18H32O2), and arachidic acid (C20:0:C20H40O2). From
Fig. 10, contents of these components were in the range of 1023%
for palmitic acid, 512% for stearic acid, and 3542% for oleic acid
with much lower amount for palmitoleic, linoleic, and arachidic
acids. It is observed that when the temperature raised from 500 to
900 8C, unsaturated fatty acid decreased while saturated fatty acids
increase. The results indicated that carbon chain length of liquid
product from pyrolysis of physic nut waste is greater than that of
diesel fuel and within the range of gas oil. It may be use on slow speed,
heavy machinery but further cracking and rening of this product are
required in order to utilize it in place of gasoline or diesel.
4. Conclusions

Fig. 9. C1C4 gaseous species produced from pyrolysis of physic nut waste.

The pyrolysis of physic nut waste was studied using TGA for
dynamic heating conditions and by fast pyrolysis for a more realistic
isothermal process. The weight loss and decomposition rate data
were used to evaluate kinetic parameters of raw material. The
kinetics of the thermal decomposition of biomass materials was
successfully modeled by a scheme consisting of three independent,
n-order parallel reactions of the main components, hemicellulose,
cellulose, and lignin. The main region of the decomposition was

162

V. Sricharoenchaikul, D. Atong / J. Anal. Appl. Pyrolysis 85 (2009) 155162

found to be between 250 and 450 8C with shift to higher temperature


at higher heating rate. Activation energy values varied between
187235 kJ/mol for cellulose, 4168 kJ/mol for hemicellulose and
97150 kJ/mol for lignin. The order of reaction is in range of 2.43.2.
Fast pyrolysis of physic nut resulted in a variety of solid charcoal,
liquid oil and gas product. Gas yield increased greatly while solid and
liquid yields decreased as temperature increases from 500 to 900 8C.
Long holding time of the volatile phase in the reactor is favorable to
gas yield. The gas products mainly consist of H2, CH4 and light
hydrocarbon. Carbon in char increased with the pyrolysis temperature while those corresponding to hydrogen and oxygen decreased
as expected. The volatile matter content of the char decreases
continuously with increasing pyrolysis temperature. Liquid oil
consisted of fatty acids such as palmitic acid, stearic acid, oleic acid
and linoleic acid. Their contents were in the range of 1023% for
palmitic acid, 512% for stearic acid, and 3542% for oleic acid with
much lower amount for palmitoleic, linoleic, and arachidic acids.
From these results, pyrolysis of physic nut waste may be considered
as an appropriate option for conversion of this residue to various
value-added products as well as fuels.
Acknowledgements
This research was carried out under the research program of the
National Metal and Material Technology Center (MT-B-49-END07-007-I). Authors would like to thank National Center of
Excellence for Environmental and Hazardous Waste Management
(NCE-EHWM) for thermogravimetric analysis.

References
[1] J. Werther, M. Saenger, E.-U. Hartge, T. Ogada, Z. Siagi, Prog. Energy Combust. Sci.
26 (2000) 1.
[2] H. Yang, R. Yan, T. Chin, D.T. Liang, H. Chen, C. Zheng, Energy Fuel 18 (2004) 1814.
[3] R. Yan, H. Yang, T. Chin, D.T. Liang, H. Chen, C. Zheng, Combust. Flame 142 (2005)
24.
[4] W. Iwasaki, Int. J. Hydrogen Energy 28 (2003) 939.
[5] G. Maschio, C. Koufopanos, A. Lucchesi, Bioresource Technol. 42 (1992) 219.
[6] A. Demirbas, Energy Convers. Manage. 43 (2002) 897.
[7] M. Garc`a-Pe`rez, A. Chaala, J. Yang, C. Roy, Fuel 80 (2001) 1245.
[8] K.G. Mansaray, A.E. Ghaly, Energy Source 21 (1999) 773.
[9] H. Teng, H.-C. Lin, J.-A. Ho, Ind. Eng. Chem. Res. 36 (1997) 3974.
[10] K. Raveendran, A. Ganesh, K.C. Khilar, Fuel 75 (1996) 987.
[11] E. Biagini, F. Lippi, L. Petarca, L. Tognotti, Fuel 81 (2002) 1041.
[12] H.B. Vuthaluru, Fuel Process. Technol. 85 (2003) 141.
[13] M.M. Nassar, Energy Source 21 (1999) 131.
[14] C.D. Blasi, G. Signorelli, C.D. Russo, G. Rea, Ind. Eng. Chem. Res. 38 (1999)
2216.
[15] R. Font, A. Marcilla, E. Verdu, J. Devesa, J. Anal. Appl. Pyrol. 21 (1991) 249.
[16] J.A. Caballero, J.A. Conesa, R. Font, A. Marcilla, J. Anal. Appl. Pyrol. 42 (1997) 159.
[17] P.T. Williams, N. Nugranad, Energy 25 (2000) 493.
[18] C. Ververis, K. Georghiou, N. Christodoulakis, P. Santas, R. Santas, Ind. Crop Prod.
19 (2004) 245.
[19] T. Fisher, M. Hajaligol, B. Waymack, D. Kellogg, J. Anal. Appl. Pyrol. 62 (2002) 331.
[20] J.J. Manya`, E. Velo, L. Puigjaner, Ind. Eng. Chem. Res. 42 (2003) 434.
[21] S. Hu, A. Jess, M. Xu, Fuel 86 (2007) 2778.
[22] R. Miranda, C. Sosa-Blanco, D. Bustos-Martnez, C. Vasile, J. Anal. Appl. Pyrol. 80
(2007) 489.
zcan, E. Putun, J. Anal. Appl. Pyrol. 52 (1999) 33.
[23] A.E. Putun, A. O
. Onay, S.H. Beis, O
.M. Kockar, J. Anal. Appl. Pyrol. 5859 (2001) 995.
[24] O
[25] H.F. Gercel, Biomass Bioenergy 23 (2002) 307.
[26] R. Zanzi, K. Sjostrom, E. Bjornbom, Biomass Bioenergy 23 (2002) 357.
[27] E. Schroder, J. Anal. Appl. Pyrol. 71 (2004) 669.
[28] S. Wang, M. Fang, C. Yu, Z. Luo, K. Cen, China Particuol. 3 (2005) 136.
[29] A. Caglar, A. Demirbas, Energy Convers. Manage. 43 (2002) 109.
[30] A. Caglar, A. Demirbas, Energy Convers. Manage. 43 (2002) 489.

Você também pode gostar