You are on page 1of 15

Journal of Environmental Chemical Engineering 8 (2020) 104561

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

Characteristics of pyrolysis products from pyrolysis and co-pyrolysis of


rubber wood and oil palm trunk biomass for biofuel and
value-added applications
Patipan Sakulkit a, Arkom Palamanit a, b, *, Racha Dejchanchaiwong c, Prasert Reubroycharoen d
a
Energy Technology Program, Faculty of Engineering, Prince of Songkla University, 15 Karnjanavanich Rd., Hat Yai, Songkhla, 90110, Thailand
b
Interdisciplinary Graduate School of Energy Systems, Prince of Songkla University, 15 Karnjanavanich Rd., Hat Yai, Songkhla, 90110, Thailand
c
Department of Chemical Engineering, Faculty of Engineering, Prince of Songkla University, 15 Karnjanavanich Rd., Hat Yai, Songkhla, 90110, Thailand
d
Department of Chemical Technology, Faculty of Science, Chulalongkorn University, Phyathai Rd., Pathumwan, Bangkok, 10330, Thailand

A R T I C L E I N F O A B S T R A C T

Editor: Yunho Lee This study investigated the yields and characteristics of bio-oil, biochar and pyrolysis gas obtained from pyrolysis
and co-pyrolysis of rubber wood sawdust (RWS) and oil palm trunk (OPT) by using an agitated bed pyrolysis
Keywords: reactor. Co-pyrolysis of RWS and OPT was performed at the mixing ratio of 50:50 (wt.). The biomass samples
Biochar were pyrolyzed at temperatures of 400, 450, and 500 ◦ C under specific conditions. The results indicated that the
Biomass
yields of the bio-oil, biochar and pyrolysis gas were in the ranges of 38.5-46.5, 22.27-28.68 and 30.1-36.9 wt.%,
Bio-oil
respectively. Pyrolysis of RWS at 500 ◦ C provided the most bio-oil. Co-pyrolysis of RWS and OPT could improve
Co-pyrolysis
Pyrolysis the product yield and quality. The bio-oil had relatively high water content, while its pH was low. The main
Pyrolysis products compounds of bio-oil were oxygenated compounds such as acetic acid, phenols and 2-Propanone, 1-hydroxy-,
determined by GC-MS. The higher heating value (HHV) of bio-oil ranged from 16.1 to 20.36 MJ/kg. The biochar
had high carbon content and low oxygen content in proximate and ultimate analysis, and based on FTIR. The
HHV of biochar was in the range from 26.35 to 29.6 MJ/kg. The biochar also was highly porous, which was
revealed by SEM and BET. The pyrolysis gas mainly contained CO, CO2, H2 and CH4, and its heating value was in
the range from 3.23 to 6.7 MJ/m3. Based on these results, the RWS and OPT are alternative and challenging
biomasses for conversion to biofuels and value-added products via pyrolysis.

1. Introduction applications in many countries, both developed and developing coun­


tries. Previously, biomass has been applied to generate heat, power, as
The applications of organic wastes and biomasses for biofuels, bio­ well as to produce biofuels (solid, gaseous and liquid biofuels).
energy or biorefining are gaining interest due to the concerns of Currently, biomass is considered a renewable alternative resource for
depleting fossil fuel resources and their environmental impacts [1,2]. energy and biochemical products in many countries, including Thailand
The heavy reliance on fossil resources not only affects security and [3,4].
environment, but it also influences sustainability and economics, In Thailand, biomass plays an important role in bioenergy and bio­
particularly in countries that need to import the fossil fuels. Thus, uti­ economy systems, because this is an agro-industrial country, with high
lization of organic wastes and biomass for biofuels, bioenergy and bio­ potential biomass resources. The main sources of biomass in Thailand
refining based on the potential of each area, region or country is are from growing, harvesting, processing, as well as replanting of crops
motivated [3]. This is due to the fact that the potential of biomass by and plants. The rubber wood and oil palm biomasses are widely avail­
region or country is different, depending on many factors such as loca­ able in the southern region of Thailand. In 2018, the plantation areas of
tion, weather, as well as agricultural and industrial activities. The rubber and oil palm trees in Thailand were 3.66 and 0.88 million
biomass has high potential for biofuels, bioenergy and biochemical hectares, respectively [5,6]. The replantation of rubber trees generates a

* Corresponding author at: Energy Technology Program, Faculty of Engineering, Prince of Songkla University, 15 Karnjanavanich Rd., Hat Yai, Songkhla, 90110,
Thailand.
E-mail address: arkom.p@psu.ac.th (A. Palamanit).

https://doi.org/10.1016/j.jece.2020.104561
Received 23 May 2020; Received in revised form 25 September 2020; Accepted 27 September 2020
Available online 1 October 2020
2213-3437/© 2020 Elsevier Ltd. All rights reserved.
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

big source of wood and biomass. The previous studies reported that pyrolysis products not only depend on types of pyrolysis reactor and
replanting of rubber trees and processing of rubber wood generated a lot pyrolysis conditions, but they are also influenced by biomass properties
of rubber wood biomass [7]. The rubber tree roots, rubber tree branches, and compositions [4].
rubber wood slabs and rubber wood sawdust are examples of rubber Applications of pyrolysis to produce bio-oil and biochar from various
wood biomass. The higher heating value (HHV) of rubber wood biomass single species biomasses of both woody and non-woody types, such as
is in the range 17.1-22.25 MJ/kg [7,8]. Some of these biomasses have corn cobs, wheat straws, rice straws and rice husks [15] oil palm
been applied in heat and power generation. However, the rubber wood biomass [16], rubber wood [8] and wood sawdust [4] have been re­
biomass still has high potential for biofuels, bioenergy and biochemical ported. However, there are not many studies related to the pyrolysis of
applications. The oil palm biomass is non-wood biomass and the main rubber wood biomass or oil palm trunks for producing bio-oil, biochar
sources of this biomass are from processing of fresh fruit bunches and and pyrolysis gas [3,10,11]. Besides pyrolysis, the co-pyrolysis of
replanting old oil palm trees. The processing of fresh fruit bunches lignocellulosic biomasses is challenging and of increasing interest.
provides many residues and wastes, including empty fruit bunches, oil Co-pyrolysis not only helps increase the utilization of low potential and
palm pressed fiber, oil palm kernel shell, decanter cake, waste water and low grade feedstocks, but it may also help improve the yield and quality
sewage sludge [3,9]. The replantation of old oil palm trees generates a of pyrolysis products as compared to pyrolysis of a single low-grade
huge source of oil palm fronds, oil palm leaves, oil palm trunks and oil feedstock. Previously, most studies of blended biomass pyrolysis have
palm roots [10]. The empty fruit bunches, oil palm pressed fiber and oil mixed biomass either with coal, sewage sludge or plastic wastes [17].
palm kernel shells have been used as solid fuel for heat and power There are very few or no studies on the co-pyrolysis of woody and
generation in industries and power plants. The higher heating value of non-woody biomasses, such as rubber wood biomass and oil palm trunk.
mentioned oil palm biomass is in the range from 12.8 to 18.7 MJ/kg [3, Therefore, the aim of this study was to investigate the yield and quality
11]. However, there is still only little applications of oil palm fronds, oil of pyrolysis and co-pyrolysis products, including bio-oil, biochar and
palm leaves, oil palm trunks and oil palm roots for biofuels, bioenergy pyrolysis gas obtained from oil palm trunk (OPT), rubber wood sawdust
and biochemical products [9]. Thus, these biomasses have high poten­ (RWS) and OPT mixed with RWS (OPT:RWS) by using an agitated py­
tial for biofuels, bioenergy and biorefinery applications. rolysis reactor.
The applications of biomass for biofuels and biorefinery are chal­
lenging because the choice of conversion process among the options 2. Materials and methods
depends on many factors such as utilization target, conversion cost,
available technology, storage and transportation, as well as biomass 2.1. Biomass preparation
properties and composition [3,10,12]. The properties and composition
of biomass are the main factors that directly influence to the quantity The biomass used in this study included OPT, RWS and OPT mixed
and quality of the obtained products [3]. Most biomasses obtained from with RWS (OPT:RWS). The fresh OPT was obtained from an old oil palm
woods, plants and crops are lignocellulosic with cellulose, hemicellulose tree harvested at 22 years age. The oil palm tree was grown at Klong
and lignin. The lignocellulosic biomass is gaining interested as a Thom District, Krabi Province, Thailand. The RWS was obtained from
promising feedstock for biofuels, energy and biorefinery applications the rubber wood processing factory located at Khlong Ngae, Sadao
[10,12]. This is due to the fact that growth of biomass absorbs carbon District, Songkhla province, Thailand. The fresh OPT was prepared by
dioxide (CO2) that is released from biomass or biofuels combustion, thus reducing the size using the saw and chopping machine (MCH-420,
utilization of bioenergy is considered carbon neutral [2,12]. The ligno­ Machinery 789, Thailand). The chopped OPT and RWS were dried with
cellulosic biomass can be classified as woody or non-woody depending solar greenhouse dryer to reduce the moisture content below 10 wt.%
on its origin. The contents of cellulose, hemicellulose and lignin in (wet basis). The dried chopped OPT and RWS was then ground with
woody and non-woody biomasses are normally different due to many grinding machine (YPS-102, Bonny, Thailand) to reduce the particle
factors such as variety, type, age, weather and growth conditions [13, size. The ground biomass samples had particle size smaller than 2 mm.
14]. The conversion of lignocellulosic biomass into biofuels, bioenergy The mixed OPT and RWS was prepared by mixing of OPT and RWS with
and biochemical products can be performed by several processes such as the ratio of 50:50 (wt.). The prepared biomass samples were kept in
mechanical, biological, thermochemical, and combined processes [10, sealed plastic bags prior to the experiments.
12].
Pyrolysis is one of alternative routes to produce the biofuels and 2.2. Determination of biomass properties and compositions
bioproducts for replacing the fossil resources [4,10]. It is widely applied
to many lignocellulosic biomasses such as wood from forestry, agricul­ 2.2.1. Particle size distribution and bulk density of ground biomass samples
tural residues and industrial wastes. Pyrolysis is the thermal decompo­ The particle size distributions of ground biomass samples were
sition of raw materials or biomass under the absence of oxygen or air at determined by using a vibratory sieve shaker (Analysett 3 pro, Fritsch,
elevated temperature higher than 400 ◦ C. The main products obtained Germany). It was equipped with seven sieve sizes, including 0.08, 0.11,
from pyrolysis include bio-oil, biochar and pyrolysis gas [4,10,15]. 0.15, 0.21, 0.30, 0.60, 1.18 mm and a bottom pan (<0.08 mm). The
There are normally two types of pyrolysis processes, including slow determination was performed following the sieve standard of ASTM E11
pyrolysis and fast pyrolysis, which are classified by heating rate, py­ [3,18]. It was done in triplicate and the result are presented as weight
rolysis temperature and vapor residence time. These pyrolysis processes percentages (wt.%). The bulk densities of the ground biomass samples
provide different yields and quality of pyrolysis products. The bio-oil were determined based on the mass and volume ratio using a modifi­
obtained from pyrolysis is a complex mixture of aromatics, carbohy­ cation of the method of Basu [19]. The ground biomass sample was filled
drate and oxygenated compounds, which are generated from the frag­ into the cylindrical container with 1000 mL volume. Then, the mass of
mentation of cellulose, hemicellulose and lignin [15]. The bio-oil with the ground biomass sample was weighted and the bulk density of
appropriate water content can be used directly as fuel for boiler or biomass was determined as the ratio of mass and volume (kg/m3) [3,
upgraded as liquid fuel for engine, as well as refined to value-added 11].
biochemical products. The biochar from pyrolysis has high carbon
content, leading to high energy content, so it is suitable for use as solid 2.2.2. Proximate and ultimate analysis
fuel for heat and power generation. The biochar also can be applied as a The gross compositions in terms of moisture content, volatile matter,
sorbent for contaminant management in soil and water [10]. The py­ fixed carbon content and ash content of biomass were determined by
rolysis gas with appropriate heating value can be used as gas fuel for proximate analysis. The analysis was performed by using the macro
heating pyrolysis or other processes. The quantity and quality of thermogravimetric analyzer (TGA 701, LECO, USA) following the

2
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Fig. 1. Schematic diagram of the pyrolysis reactor.

procedure of ASTM D7582 [3,10]. The determination was carried out in


triplicate and results are presented as mean value with standard devia­ Cellulose = ADF-ADL (wt.%, dry basis) (1)
tion, on as received basis (wt.%, as received basis). The elemental
compositions of biomass samples were determined by ultimate analysis Hemicellulose = NDF-ADF (wt.%, dry basis) (2)
using CHNS/O Analyzer (Flash 2000, Thermo Scientific, Italy), Lignin = ADL (wt.%, dry basis) (3)
following the ASTM D4239 and EN 15104 standard. The determination
was performed in triplicate and the mean value with standard deviation Extractives = 100- Cellulose - Hemicellulose - Lignin (wt.%, dry basis) (4)
are reported for contents of carbon (C), hydrogen (H), nitrogen (N),
oxygen (O) and sulphur (S) based on dry basis (wt.%, dry basis) [3,10].

2.2.3. Energy content and potential use as energy equivalent with fossil 2.3. Experimental set-up
fuels
The energy content or heating value of biomass samples is reported The schematic diagram of agitated pyrolysis reactor used in this
in terms of higher heating value (HHV). The HHV of biomass was study is shown in Fig. 1 (modified from Palamanit et al. [10]). The main
determined directly by using bomb calorimeter (IKA C5000, Germany) components of this reactor include a reaction chamber, heater with
according to EN 14918 procedure. The HHV of biomass samples ob­ controller, high torque direct current motor, agitated blade connected to
tained from bomb calorimeter was then converted to the potential en­ a magnetic coupling drive, motor speed controller, hopper and control
ergy equivalent with fossil fuels, including coal (bituminous), crude oil, valve, cyclone, shell and tube condensers, water cooling bath, nitrogen
natural gas (NG) and liquefied petroleum gas (LPG). The reference HHV gas tank with regulation system, nitrogen gas flow meter, and flow
of coal, crude oil, NG and LPG was 28.87 MJ/kg, 38.78 MJ/L, 38.95 MJ/ meter for cold water. The shape of reactor chamber was cylindrical and
m3 and 50.08 MJ/kg, respectively [3,20,21]. it was made from stainless steel with a diameter, height and wall
thickness of 13,15 and 0.5 cm, respectively. The reaction chamber was
2.2.4. Thermogravimetric analysis heated by electrical heater with the power of 3 kW. The temperature of
The thermal decomposition behavior of biomass sample under inert biomass sample during pyrolyzing can be measured by K-type thermo­
condition was observed via thermogravimetric analysis (TGA) and de­ couple equipped with stainless steel shield. The PID-controller was used
rivative thermogravimetric analysis (DTG) using the thermogravimetric to control the temperature of biomass. The biomass sample was fed into
analyzer (Perkin Elmer, USA), according to ASTM E1131 procedure. The reaction chamber via hopper equipped with control valve. During py­
biomass sample of 10 mg was subjected to temperature ranging from 50 rolyzing, the biomass sample inside reaction chamber was mixed by

C to 1000 ◦ C, at heating rate of 10 ◦ C/min under a nitrogen (N2) gas mixing blade connected to electrical motor (4RK25GN-C, Zhengke
atmosphere [3,10]. Electromotor Co., Ltd., China) via magnetic coupling drive connector.
The speed of agitated blade can be controlled by adjusting the motor
2.2.5. Lignocellulosic composition shaft speed using the controller. The nitrogen (N2) gas was supplied to
The biomass samples were determined for chemical compositions in the reactor through the bottom of the reactor. The flow rate of N2 gas can
terms of cellulose, hemicellulose and lignin contents. These contents be controlled by controlled valve and its flow rate was measured by flow
were determined via the method developed by Georing and Van Soest meter (RMA-21-SSV, Dwyer, USA). The vapor of pyrolyzed biomass
[22], Van Soest [23], Reza et al. [24], and Kambo and Dutta [25]. Ac­ exited from the top of reactor and flowed to the cyclone, which was used
cording to this method, the acid detergent fiber (ADF), acid detergent to remove the solid particles from pyrolysis vapor. The vapor was
lignin (ADL) and neutral detergent fiber (NDF) were determined, and condensed in three shell and tube condensers. The bio-oil was collected
then the percentages of cellulose, hemicellulose and lignin were calcu­ at collectors of condensers. Cold water was supplied to each condenser
lated using the equations (1)–(3). The part of extractives was calculated from cooling water bath (CBD-30-P, Scientific Promotion Co., Ltd,
by difference method as shown in equation (4). All results are presented Thailand). The flow rate of cold water can be controlled by balancing
as mean value and standard deviation on dry basis (wt.%, dry basis) [3, valve and measured by rotameter. The incondensable vapor could be
10,11]. recycled to use as carrier gas by control valve or released to environment

3
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

through the wet scrubber. analysis using the same method described in previous section. The en­
ergy content of bio-oil was considered in terms of higher heating value
2.4. Experimental procedure (HHV). It was determined by a bomb calorimeter using the same method
described in previous section. Before determining the HHV of bio-oil, the
The experiments of this study were of batch type runs. For each batch sample of bio-oil was dried in a hot air in oven at 60 ◦ C until its weight
experiment, 150 g of biomass was pyrolyzed. The experiment was was constant [10]. The Van Krevelen diagram (H/C and O/C atomic
started by heating the reaction chamber to a temperature of 97 ◦ C, then ratios) of water-free bio-oil (dry basis) was also compared to conven­
the biomass was fed into the reaction chamber via the hopper and tional oils, including crude oil and heavy fuel oil [29,30].
control valve. The valve was closed and the biomass sample was heated
continuously to desired temperature (400, 450 and 500 ◦ C). These 2.5.4. Chemical compounds
temperatures were selected based on the thermal decomposition The chemical compounds of the bio-oil were determined by using gas
behavior of biomass samples, which were indicated by TGA and DTG chromatography-mass spectroscopy (GC-MS) (Agilent, USA). The
analyses. The results showed that most thermal decomposition of GC–MS system was equipped with the column of Agilent CP9205 VF-
biomass occurred at temperatures ranging from 200 to 400 ◦ C. Thus, the WAXms. The specification of column was 30 m × 250 μm × 0.25 μm
pyrolysis needs to have temperature above 400 ◦ C to make sure that the (length × ID × film thickness). The stationary phase was polyethylene
hemicellulose and cellulose are mostly decomposed. The pyrolysis glycol. The helium gas was used as carrier gas at the flow rate of 1 mL/
temperatures were also referred from the previous study of Palamanit min. The injection port and detector were both operated at temperature
et al. [10]. During heating, the biomass was continuously mixed by of 250 ◦ C. The GC oven was heated to 70 ◦ C for 2 min, and then it was
mixing blade at the speed of 10 rpm. When the biomass temperature was heated to the temperature of 250 ◦ C for 10 min at a heating rate of 5 ◦ C/
increased to 250 ◦ C, the N2 gas with the flow rate of 2 L/min was purged min. The injection was performed with 10 μl of bio-oil sample. The
into the reaction chamber to maintain an inert atmosphere inside the chemical compounds of bio-oil sample were then identified based on the
reactor. When the biomass temperature reached the set point tempera­ peak matching of the mass spectra from the National Institute of Stan­
ture, the N2 gas flow rate was increased to 5 L/min. The vapor produced dards and Technology (NIST) mass spectral library [14,31,32].
during pyrolyzing was immediately condensed in condensers with cold
water at temperature of 10 ◦ C and a flow rate of 8 L/min. The biomass 2.6. Characteristics of biochar
sample was pyrolyzed at set point temperature for 30 min for each
experiment. The condensable and incondensable vapors were consid­ 2.6.1. Particle size distribution, proximate analysis, ultimate analysis and
ered bio-oil and pyrolysis gas, respectively. The solid residue remaining energy content
in reaction chamber was considered biochar. The yields of bio-oil and The particle size distribution, proximate analysis and ultimate
biochar were directly determined by weighing. The weight of pyrolysis analysis of biochar were determined by using the same method
gas was determined by the difference method. The sum of all pyrolysis described in the previous sections. The energy content of biochar was
products was considered 100 wt.% based on the mass balance of whole reported as HHV. The energy content of biochar was reported as HHV.
pyrolysis process. The investigated properties and compositions of They were determined by calculating with the equations (eq. 5) [33]:
bio-oil, biochar and pyrolysis gas were then determined and analyzed. HHV = 0.341C + 1.322H – 0.12O – 0.12 N + 0.0686S – 0.0153Ash
(MJ/kg) (5)
2.5. Characteristics of bio-oil where, C, H, O, S and Ash are the components of biomass obtained
from ultimate analysis and proximate analysis [33]. The Van Krevelen
2.5.1. Water content, pH value and density diagram (H/C and O/C atomic ratios) of biochar was compared between
The water content of the bio-oil was determined by volumetric Karl- biomass and coal [34].
Fischer titration technique using a Mettler Toledo V20 automatic titrator
[10,26]. The pH value of bio-oil was measured by a pH meter (UB-10 2.6.2. Surface morphology
Denver Instrument) at room temperature. Before measurement, the in­ The scanning electron microscope (SEM) (JSM-5800 LV, JEOL,
strument was calibrated with standard liquid with pH of 4, 7, and 10 Japan) was used to observe the surface morphology of the biomass and
[10,27,28]. The density of bio-oil was determined based on the weight biochar sample. Prior to observation with SEM, the sample was coated
and volume ratio using a density bottle at room temperature [16,26]. with gold by sputtering method. The SEM was performed under high
The determination was performed in triplicate and the result was re­ vacuum condition with accelerating voltage of 20 kV, and a secondary
ported as mean value with standard deviation value. electron (SE2) detector with magnification of 500X [10].

2.5.2. Solid and ash content 2.6.3. Surface area and average pore diameter
The solid content was determined by vacuum filtration technique. The surface area of the biomass and biochar sample was determined
The bio-oil with 3 g was filtered through a pre-dried Whatman No. 2 by Brunauer– Emmett–Teller (BET) method (ASAP2460, Micromeritics,
filter paper. The filter paper was then washed with of 99.9 % ethanol USA) using the static volumetric N2 gas adsorption method. The sample
until the filtrate was clear to ensure that there was no bio-oil left on the was pretreated under vacuum condition at temperature of 80-200 ◦ C for
filter paper. The filter paper with the solid residue was pre-dried with 15 hours. The adsorption-desorption isotherm was determined by the
ambient air for 15 min and then it was dried with hot air in an oven at static volumetric method. The N2 gas was used as carrier gas and
temperature at 105 ◦ C for 30 min. The dried filter paper was cooled in adsorbed gas. The pore volume was obtained from the adsorption
the desiccator prior to weigh and the solid content was calculated and isotherm with the multi condensation point (p/p0 = 0.05-0.03, 10-20
reported as mass ratio (wt.%) [10,27,28]. The ash content of bio-oil was points). The pore structure was determined from the adsorption
determined by burning method. The sample of bio-oil was burned in a isotherm and the average width was calculated by using the formula 4
furnace at temperature of 775 ◦ C with air for 24 hours. The solid residue V/A (“V” represents the pore volume while “A” denotes the adsorbed bet
remaining in crucible was considered as ash [16,27]. The determination specific surface area) [10].
was performed in triplicate and the result was reported as mean value
with standard deviation value. 2.6.4. Functional groups
The Fourier Transform Infrared Spectroscopy (FTIR) technique was
2.5.3. Ultimate analysis and energy content used to characterize the organic functional groups of biochar. In this
The elemental composition of bio-oil was determined by ultimate study, the biochar obtained from pyrolysis temperature of 500 ◦ C was

4
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 1
Bulk density, and proximate, and ultimate analysis results of ground biomass
samples
Biomass type
Characteristic
OPT RWS OPT:RWS

Proximate analysis (wt.%, as received basis)


Moisture content 9.65 ± 0.03 4.95 ± 0.02 7.3 ± 0.03
Volatile matter 74.1 ± 0.3 77.65 ± 75.87 ±
0.36 0.33
Fixed carbon 13.96 ± 15.3 ± 0.47 14.63 ±
0.31 0.39
Ash content 2.3 ± 0.03 2.1 ± 0.18 2.2 ± 0.11
Ultimate analysis (wt.%, dry basis)
Carbon 41.63 ± 45.43 ± 43.53 ±
0.17 0.04 0.11
Hydrogen 5.99 ± 0.23 6.01 ± 0.05 6.00 ± 0.14
Nitrogen 0.15 0.09 0.12
Sulphur 0.10 0.07 0.09
Oxygen 43.02 ± 41.72 ± 42.37 ± 0.2
0.12 0.28
Lignocellulosic compositions (wt.
%)
Cellulose 46.24 ± 56.6 ± 0.8 51.42 ±
Fig. 2. Particle size distributions of ground biomass samples.
0.04 0.42
Hemicellulose 20.84 ± 17.4 ± 0.26 19.12 ±
analyzed for functional groups by the FTIR (Vertex70, Bruker, Ger­ 0.12 0.19
Lignin 4.95 ± 0.08 17.56 ± 11.25 ± 0.3
many). Each spectrum was an average of 32 scans from 400-4000 cm-1 at
0.52
4 cm-1 spectral resolution. The FTIR data interpretation of each spec­ Extractive 27.97 ± 8.45 ± 0.18 18.21 ±
trum was based on a literature survey using extensive elucidation. 0.38 0.28
Higher heating value (MJ/kg) 16.31 ± 17.7 ± 0.09 16.85 ±
0.05 0.02
2.7. Pyrolysis gas compositions and heating values Bulk density (kg/m3) 189 ± 0.27 290.4 ± 267.7 ±
0.45 0.56
A sample of gas obtained from pyrolysis temperature of 500 ◦ C was
collected in gas bag. Then, composition of the gas was determined
rate of carrier gas had risk of decreased to bio-oil yield. The small
immediately by gas chromatography (GC) (GC-2014, Shimadzu). This
biomass particles are easily carried out from the reaction zone before
GC was equipped with a thermal conductivity detector (TCD). The col­
pyrolysis reaction is complete. The loss of small biomass particles also
umn of CG was a 2.0 m packed column (Shin carbon ST 100/120 Restek)
led to bio-oil with a high solids content. In the case of pyrolysis of large
for indicating the concentrations (vol. %) of H2, CO2, CO, and CH4. The
biomass particles, a previous study found that the bio-oil and pyrolysis
temperature of injection port, oven and detector was 120, 50 and 100
gas yields were decreased by increased biomass particles size, while the

C, respectively. The argon gas was used as carrier gas at a flow rate of
biochar yield increased. Thus, in practice the biomass particle size for
15 mL/min. The gas sample with the volume of 0.5 mL was injected to
pyrolysis may ranges from micrometers to centimeters, depending on
GC [35]. The heating value of the pyrolysis gas was then calculated as
the reactor type and operating conditions. However, the preparation of
follows (Eq. 6) [10,36]:
biomass particles needs to consider the related aspects such as cost,
Heating value = 13.1(%CO/100) + 13.2(%H2/100) + 41.2(%CH4/100) (MJ/ energy consumption, transportation, storage, as well as the feeding
m 3) (6) system of pyrolysis reactor [3,37].
The bulk density of ground biomass samples is shown in Table 1. It is
where CO, H2, and CH4 are the percentages (vol.%) of carbon monoxide, seen that the bulk density of the biomass samples was in the range of
hydrogen, and methane in the pyrolysis gas product. 189– 290.4 kg/m3. The ground RWS had the highest bulk density (290.4
kg/m3), followed by the ground OPT:RWS (267.7 kg/m3) and ground
3. Results and discussion OPT (189 kg/m3). The bulk density of ground biomass samples was
corresponding to the biomass particles size as discussed in previous
3.1. Characteristics of raw biomass paragraph. The bulk density of ground OPT was low due to its fibrous
structure [3]. Generally, the bulk density of biomass depends on the
3.1.1. Particle size distribution and bulk density of ground biomass sample particle size, shape, moisture content, particle density and surface
Fig. 2 depicts the particle size distributions of the ground biomass characteristics. The previous studies indicate that the bulk density of
samples. The results showed that most of the ground biomass samples different biomasses, including non-woody biomass such as cereal grains
had particle sizes ranging from 0.3 to 1.18 mm. For RWS, its particle size and straws ranged from 15 to 200 kg/m3, 280 - 480 kg/m3 for woody
was in the range of 0.3-0.6 mm, which was relatively smaller than the biomass [38] and 152-896 kg/m3 for oil palm biomass [3]. As compared
ground OPT and OPT mixed with RWS (0.6-1.18 mm). The size of to the bulk density of coal, it is clear that the bulk density of lignocel­
biomass particle a physical property that influences yield and quality of lulosic biomass is lower [14]. For pyrolysis applications, the bulk density
pyrolysis products. This is because the biomass particle size affects heat of biomass particles directly affects the biomass volume present in the
transfer during pyrolyzing. Pyrolysis of biomass with small particles reaction chamber, consequently it influences the heat transfer behavior
provides high heat transfer rate because the surface area of small par­ during pyrolyzing. The bulk density of biomass particles also relates to
ticles is larger than of large biomass particles. For bio-oil production, the energy density (MJ/m3), storage area, and handling and transportation
heat transfer rate or heating rate not only affected the bio-oil yield but it costs [10,14].
also influenced the quality of bio-oil both in fixed bed and moving bed
pyrolysis reactors [26,27]. A previous study indicated that pyrolysis of
very small biomass particles with moving bed reactor or using high flow

5
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

3.1.2. Proximate and ultimate analysis results


Table 1 shows the results of proximate and ultimate analysis of the
OPT, RWS and OPT:RSW. The proximate analysis indicated that the
moisture content, volatile matter, fixed carbon content and ash content
of biomass samples were in the ranges of 4.95-9.65, 74.09-77.65, 13.96-
15.3, and 2.1-2.3 wt.%, respectively. The biomass samples had low
moisture contents (<10 wt.%), which is appropriate for pyrolysis ap­
plications [4,10]. This is because the moisture content of biomass
strongly affects the yield and quality of bio-oil [39]. Moreover, a high
moisture content of biomass also consumes extra heat for evaporating
the moisture and increasing the vapor temperature to desired set point
temperature. Thus, pyrolysis of moist biomass gives bio-oil with high
water content, and has a large thermal energy consumption for pyrolysis
reactor [40]. Regarding volatile matter, it is seen that the volatile matter
of RWS was higher than OPT and OPT:RWS. This result indicates that the
volatile matter of these biomasses was not much different from other
biomasses, for example switchgrass, macadamia shells, coconut shells,
rice straw, redwood, wheat straw, corn stover, corn cob and oil palm
biomass [3,11,14,15]. The volatile matter is the main component in
biomass for producing bio-oil by pyrolysis. The higher volatile matter in
biomass is favorable for bio-oil production with pyrolysis since the
bio-oil tends to depend on the volatile matter [41]. Considering the fixed
carbon content of biomass samples, it indicated that the RWS had higher
fixed carbon content as compared to OPT and OPT:RWS. The fixed
carbon of biomass is the component that can be converted into biochar.
Thus, the biochar yield of pyrolysis products depends on the fixed car­
bon content of that biomass. For producing the biochar, the biomass
with high fixed carbon content is appropriate. The inorganic composi­
tions of biomass are considered as ash. The biomass used in this study
had low ash content (2.1-2.3 wt.%). The ash content of RWS was lower
than OPT and OPT:RWS. Some previous studies indicated that the ash
content of woody biomass was lower than of non-woody biomass [14].
The low ash content of biomass also enhanced the quality and quantity
of bio-oil by avoiding the secondary reaction of condensable vapors
[42], as well as the reduction of solid content in bio-oil.
For the elemental composition of biomass samples, the ultimate
analysis showed that the contents of C, H, N, S, and O of OPT, RWS and
OPT:RWS were in the ranges of 41.63-45.43, 5.99-6.01, 0.09-0.15, 0.07-
0.1, and 41.72-43.02 wt.%, respectively. The contents of C and H in RSW
were higher than in OPT, while the N, S and O of RSW were lower than in Fig. 3. TGA (a) and DTG (b) of ground biomass samples.
OPT. The result of ultimate analysis was consistent with the result of
proximate analysis. As compared to the other biomasses, it indicated the
were 46.24-56.6, 20.84-17.4, 4.95-17.56, and 8.45-27.97 wt.%,
carbon and hydrogen contents of RSW and OPT were relatively high
respectively. The main component in biomass samples was cellulose.
compared to cassava rhizomes, cassava stalk, wheat straw, corn stalk,
The hemicellulose content of RWS was lower than lignin content. But,
wood sawdust, corncob, and rice husk, bamboo, oak wood and birch
the lignin content of OPT was relatively similar to the hemicellulose
wood [4,14,15,27,28]. Normally, biomass with high carbon and
content. The previous studies indicated that the biomass with high cel­
hydrogen content provides a high HHV and LHV [10,30]. In the part of
lulose and hemicellulose is favorable for producing bio-oil using pyrol­
oxygen content, it was relatively high. For pyrolysis process, the oxygen
ysis process. This is because these compositions promote high yield of
content of biomass should be low, while the carbon and hydrogen
bio-oil as compared to the use of biomass with high lignin content
contents need to be high. This is due to the fact that the carbon and
[10,44]. Pyrolysis of biomass with high lignin content usually provides
hydrogen in biomass can be converted into useful aromatic compounds
high biochar yield. This is due to the thermal decomposition of cellulose
in bio-oil. On the other hand, the oxygen will be bonded with the hy­
and hemicellulose that occurs at a lower temperature comparing to the
drocarbon molecules during pyrolysis process as oxygenated com­
lignin [10,11]. The effects of lignocellulose content of biomass on bio-oil
pounds, which reduce the quality of bio-oil. Moreover, the pyrolysis of
and biochar yield were studied by many authors. Qu et al. [45] found
biomass with high oxygen content is also risk to obtain the bio-oil with
that the pyrolysis of cellulose provided a high bio-oil yield as compared
high water content. This is because water can be formed from the re­
to pyrolysis of hemicellulose because cellulose is more volatile than
action between hydrogen and oxygen [43]. The nitrogen and sulphur
hemicellulose. Kim et al. [46] concluded that there was a higher bio-oil
content of biomass samples were low. The low nitrogen and sulphur
yield when using biomass with higher cellulose and hemicellulose
contents of biomass help avoid the partial formation of NOx and SOx
contents. Quan et al. [47] pyrolyzed cellulose, hemicellulose, and lignin
during oxidation of nitrogen and sulphur with oxygen in biomass. Most
at 500 ◦ C and they found that the pyrolysis of these components pro­
of sulphur and nitrogen remain in the biochar [4,10].
vided liquid yields of 18.67, 30.83, and 0.5 wt.%, respectively. In
addition to the bio-oil yield, variations in the composition of the cellu­
3.1.3. Lignocellulosic compositions
lose, hemicellulose, and lignin in the biomass also influence the chem­
Table 1 shows chemical composition of biomass samples in term of
ical compounds of the bio-oil or liquid product [44]. In addition, Wang
cellulose, hemicellulose, lignin and extractives. The results indicated
et al. [48] reported that the extractives in biomass could enhance the
that cellulose, hemicellulose, lignin and extractives of biomass samples

6
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 2
Commercial fuel equivalents for same energy as one kilogram of biomass
Fossil fuel amount with equivalent energy
Type of
biomass HHV (MJ/ Crude oil Coal LPG NG
kg) (L) (kg) (kg) (m3)

OPT 16.31 0.42 0.56 0.33 0.42


RWS 17.7 0.46 0.61 0.35 0.45
OPT:RWS 16.85 0.43 0.58 0.34 0.43

bio-oil yield and suppress the char and gas production when using corn
stalks and wheat straws as feedstocks for pyrolysis.

3.1.4. Thermal decomposition behaviour


The thermal decomposition behavior of biomass samples was
observed via TGA and DTG as shown in Fig. 3(a) and (b), respectively.
These figures clearly indicate that there were four main stages of ther­
mal decomposition of biomass samples. The first stage occurred at
temperatures of 50-120 ◦ C, in which the weight of biomass samples
slightly decreased. The mass reduction was less than 10% due to evap­ Fig. 4. Effects of pyrolysis temperature on product yields from OPT.
oration of the moisture present in biomass. In the second stage which
occurred at temperatures of 120-200 ◦ C, the weight of the biomass
samples was relatively constant due to less evaporation of the light
volatile compounds. The thermal energy supplied to the biomass at this
stage was used to increase its temperature. The third stage is the main
thermal decomposition of the biomass samples as clearly depicted in
Fig. 3(a) and (b). This stage occurred at temperatures of 200-400 ◦ C,
with a mass loss rate of about 65-70%. Most of the volatile organics,
including cellulose and hemicellulose were decomposed by heat to
condensable and incondensable vapors or gases, which correspond to
the thermal decomposition of general lignocellulosic biomass [3,10].
The last stage was present at temperatures above 400 ◦ C, and the ther­
mal degradation of the biomass samples was relatively slow at this stage.
The mass loss at temperatures 400-1000 ◦ C was about 10-15% due to the
decomposition of the lignin. The thermal decomposition trend of
biomass samples was consistent with the proximate analysis results and
lignocellulosic components as shown in previous sections. The previous
studies indicated that the thermal decomposition of hemicellulose, cel­
lulose and lignin of biomass occurred in temperature ranges of 250-350,
325-400 and 300-900 ◦ C, respectively [4,10,44].
Fig. 5. Effects of pyrolysis temperature on product yields from RWS.
3.1.5. Energy content and potential use as energy equivalent with fossil
fuels
The energy content and potential use as energy equivalent with fossil
fuels of biomass samples is shown in Table 1 and 2. The energy content
of biomass samples was presented as higher heating value (HHV). The
results showed that the HHV of OPT, RWS and OPT:RWS was 16.31,
17.7 and 16.85 MJ/kg, respectively. The heating value of biomass de­
pends on many factors such type, composition, age and plantation re­
gion and conditions. The composition of biomass is one of most
important factors that influences its heating value. Based on this result,
it is observed that the HHVs of biomass samples were consistent with the
proximate and ultimate analysis results. Previous studies also indicated
that the biomass with high cellulose and lignin tended to have high
HHV. Demirbaş [49] reported that the lignin has HHV better than cel­
lulose and starch. For the pyrolysis process, the heating value of the
biomass did not clearly affect the heating value of the bio-oil. This is
because different types of biomass with the same heating value may
have different compositions of volatile matter, fixed carbon content and
lignocellulose content. Thus, these components can be converted into
condensable vapors under different conditions during the pyrolysis
process [10,50]. The potential use of the biomass samples as energy
Fig. 6. Effects of pyrolysis temperature on product yields from OPT:RWS.
equivalent with crude oil, coal, LPG, and NG were 0.42-0.46 L, 0.56-0.61
kg, 0.33-0.35 kg, and 0.42-0.45 m3, respectively. It indicated that
biomass samples had relatively low potential for use as energy since the
energy content of biomass was low. However, utilization or

7
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 3 should have the most volatile matter, leading to an increase in the bio-oil
Basic properties of bio-oil yield [10]. The lower bio-oil yield of OPT is due to high ash content in
Property the biomass samples. The same trend also occurred in the study of Fahmi
Pyrolysis et al. [51]. Also, the RWS have high density and particle size was smaller
Biomass temperature Water pH Solid Ash Density
content value content content (kg/m3) than of OPT and RWS mixed with OPT, as shown in Fig. 2. The smaller
(◦ C)
(wt.%) (wt.%) (wt.%) particle size of biomass sample led to good heat transfer rate, which
2.45
strongly affects the bio-oil yield [3]. However, previous studies reported
76.36 ± 0.33 ± 0.3 ± 1006 ± that the bio-oil yield of pyrolysis not only depends on biomass compo­
400 ±
0.11 0.05 0.05 0.06
0.02 sition and properties, but is also influenced by the operating parameters
2.32 and reactor type [16,52,53]. The pyrolysis of biomass at a high heating
73.54 ± 0.64 ± 0.42 ± 1002 ±
OPT 450
0.7
±
0.04 0.05 0.16 rate using moving bed reactors provided bio-oil yield exceeding 47 (wt.
0.01
2.84 %) [16,37]. Pyrolysis of biomass with moving bed reactors normally
74.12 ± 0.82 ± 0.44 ± 1005 ±
500 ± provides higher bio-oil yield as compared to fixed bed reactors [53–54].
0.96 0.06 0.06 0.07
0.01 In contrast, pyrolysis of biomass with high fixed carbon content led
65.48 ± 0.27 ± 0.3 ± 1008 ± to high yield of biochar. In the case of biochar, the highest biochar yield
400 2.35
0.5 0.02 0.03 0.05
2.28
was obtained from RWS (28.68 wt.%) and the lowest from OPT (24.2 wt.
63.78 ± 0.31 ± 0.37 ± 1018 ± %). The biochar yield was high at lower temperature (400 ◦ C) and it was
RWS 450 ±
0.53 0.08 0.01 0.31
0.02 significantly decreased when increase in temperature [39,55]. This is
500
63.17 ± 2.4 ± 0.4 ± 0.4 ± 1024 ± due to a greater thermal decomposition of lignocellulosic components at
0.71 0.02 0.05 0.04 0.15
higher temperature [50,52]. However, the gas yield showed the oppo­
67.4 ± 0.57 ± 0.3 ± 1008 ±
400
0.65
2.66
0.06 0.09 0.46 site trend, it decreased with increasing temperature due to the con­
75.66 ± 0.64 ± 0.36 ± 1011 ± densable gases contributing to bio-oil. The highest and lowest yield for
OPT: 450 2.75
RWS
0.35 0.04 0.07 0.22 the gas were 36.9 wt.% and 30.1 wt.% obtained from OPT and RWS at
2.58 500 ◦ C, respectively.
71.5 ± 0.74 ± 0.4 ± 1019 ±
500 ±
0.81 0.08 0.01 0.22
0.02
3.3. Properties of bio-oil

co-utilization of biomass as bioenergy can help reduce the reliance on 3.3.1. Water content
fossil fuels and environmental impacts. Table 3 presents the water contents of bio-oil obtained from pyrolysis
of OPT, RWS and OPT:RWS at different pyrolysis temperatures. The
results showed that the water content of the bio-oil was in the range of
3.2. Distribution of pyrolysis product yields 63.17-76.36 wt.%. These results indicated that pyrolysis of OPT, RWS
and OPT:RWS at different pyrolysis temperatures had a small effect on
Fig. 4–6 depict the product yields obtained from the pyrolysis of the water content of the bio-oil. The water content in the bio-oil was
OPT, RWS, and OPT:RWS at temperatures of 400, 450, and 500 ◦ C. The consistent with the moisture content of raw biomass samples, as well as
results showed that the product yields of the bio-oil, biochar and py­ their oxygen and hydrogen contents. The bio-oil obtained from pyrolysis
rolysis gas were in the ranges of 38.5-46.5, 22.27-28.68 and 30.1-36.9 of RWS contained less water than that from OPT and OPT:RWS. This is
wt.%, respectively. These results indicated that pyrolysis temperature because the moisture content of RWS was lower than of other biomass
and biomass type affected the product yields. As can be seen in that samples. The previous studies reported that the water content of bio-oil
Figures, an increase in pyrolysis temperature from 400 ◦ C to 500 ◦ C led is depended on the moisture content of raw biomass and is also influ­
to obtain more bio-oil for all biomass samples. The higher bio-oil led to enced by the oxygen content in biomass [55,56]. The pyrolysis of
lower biochar and pyrolysis gas yield. Considering the effect of pyrolysis biomass with a high oxygen content has risk of producing bio-oil with
temperature, it is clearly seen that pyrolysis at temperature of 500 ◦ C high water content. This is because water can be formed from the re­
provided the highest bio-oil yield, while the biochar was highest at action between hydrogen and oxygen of biomass [43]. The results ob­
temperature of 400 ◦ C for all biomass samples. An increase in pyrolysis tained from this study indicated that the water content of the bio-oil was
temperature provided the higher bio-oil yield due to more thermal higher than 50 wt.%. The high water content of bio-oil is due to the used
decomposition of lignocellulosic components, particularly the hemicel­ pyrolysis reactor providing a low heating rate, which was about 14
lulose and cellulose as discussed in section 3.1.4. In the case of biomass ◦
C/min. However, the water content of bio-oil obtained from this study
types, the results indicated that pyrolysis of RWS obtained the highest was relatively similar with many previous studies. Chen et al. [57]
liquid yield and co-pyrolysis of OPT with RWS helped to improve the report that the water content of the obtained bio-oil was 64.4-75.1 wt.%.
bio-oil yield as compared to pyrolysis of OPT only. However, it is Gómez et al. [58] found that the slow pyrolysis of biomass provided
observed that at each pyrolysis temperature, the biomass type had bio-oil with water content higher than 50 wt.%. Recently, Palamanit
smaller effect on the products yields than pyrolysis temperature. This is et al. [10] showed that the slow pyrolysis of oil palm biomass provided
because the compositions of biomass samples are not much different as liquid products with the water content ranging from 61.1 to 68.3 wt.%.
indicated by volatile matter, fixed carbon content and lignocellulosic In the case of fast pyrolysis, the water content of the bio-oil usually
composition. Based on these compositions, the yields of bio-oil and varies in the range 10-40 wt.% (wet basis), depending on the biomass
biochar products were consistent with the proximate analysis results and type and composition, as well as the pyrolysis conditions [10,59].
the lignocellulosic components of the biomass samples. The previous Moreover, the water content in bio-oil not only depends on the moisture
studies revealed that pyrolysis of the biomass with high volatile matter content and oxygen content of raw biomass but is also influenced by
and low ash content resulted in a high amount of bio-oil [10,27]. The pyrolysis conditions such as heating rate, reactor type, vapor residence
results of the pyrolysis products also correspond to the lignocellulose time and pyrolysis temperature [10,59]. The water content in bio-oil has
content of the biomass samples. As can be seen in Table 1, the RWS both positive and negative impacts. The bio-oil with low water content
contained high hemicellulose and cellulose content, while the lignin and provides high HHV and good stability, while the high water content in
extractives were low. Thus, the pyrolysis of the RWS provided the bio-oil helps improve the flow ability [26,27].
highest amount of bio-oil. Qu et al. [45] reported that the cellulose is
more volatile, therefore, the RWS with the largest amount of cellulose

8
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 4
Ultimate analysis results and higher heating value of bio-oil
Ultimate analysis (wt.%, as received basis)
Biomass Pyrolysis temperature (◦ C) HHV (MJ/kg)
Carbon Hydrogen Nitrogen Sulphur *Oxygen

400 ◦
C 13.8 ± 0.2 9.91 ± 0.19 0.07 <0.01 76.21 ± 0.39 17.3 ± 0.06
OPT 450 ◦
C 14.63 ± 0.3 9.92 ± 0.21 0.08 <0.01 75.36 ± 0.24 19.51 ± 0.08
500 ◦
C 15.81 ± 0.21 9.44 ± 0.17 0.16 <0.01 74.58 ± 0.37 20.3±0.34
400 ◦
C 16.93 ± 0.18 9.83 ± 0.23 0.59 ± 0.18 <0.01 72.64 ± 0.41 18.07 ± 0.26
RWS 450 ◦
C 20.97 ± 0.13 9.49 ± 0.12 0.10 ± 0.01 <0.01 69.43 ± 0.24 20.33 ± 0.15
500 ◦
C 21.53 ± 0.36 9.22 ± 0.26 0.23 ± 0.01 <0.01 69.01 ± 0.24 20.36±0.08
400 ◦
C 17.62 ± 0.25 9.74 ± 0.07 0.16 ± 0.01 <0.01 72.47 ± 0.28 16.1 ± 0.34
OPT:RWS 450 ◦
C 18.33 ± 0.08 9.90 ± 0.21 0.70 ± 0.01 <0.01 71.06 ± 0.3 18.48 ± 0.05
500 ◦
C 22.62 ± 0.1 9.12 ± 0.18 0.18 ± 0.01 <0.01 68.07 ± 0.09 19.24 ± 0.1
*
Oxygen by difference.

3.3.2. pH, solids content, ash content and density


The pH, solids content, ash content and density of bio-oil are also
shown in Table 3. It is seen that the pH, solids content, ash content and
density of bio-oil were in the ranges 2.28-2.84, 0.27-0.82 wt.%, 0.3-0.44
wt.%) and 1002-1024 kg/m3, respectively. These results imply that
biomass type and pyrolysis temperature had small effects on these
properties of bio-oil. The pH of bio-oil obtained from this study was
similar to pH of the bio-oil from previous studies [10,26]. This is because
bio-oil contains a high amount of organic acids, including acetic acid
and formic acid [16,35]. Other groups of chemical compounds in bio-oil
that influence the acidity include phenolics and fatty, resin and hydroxy
acids. Recently, Chorazy et al. [60] reported that the pH of bio-oil from
lignocellulosic biomass was in the range 2-3. The presence of acids in
bio-oil was due to the degradation of hemicellulose and cellulose of
biomass, which has high oxygen content. The high acid value of bio-oil is
a disadvantage of this product for direct use or applications, since it can
cause corrosion of materials during storage and applications [60].
In the part of solids content, it is seen that biomass types and py­
rolysis temperature had small effect on the solids content of bio-oil. The
solids content of the bio-oil samples obtained from this work was rela­ Fig. 7. Van Krevelen diagram (H/C and O/C atomic ratios) for water-free bio-
tively low as compared to other types of moving bed reactors such as oil (500 ◦ C) and conventional oils.
fluidized bed and twin screw reactors [26,28]. The solids content in
bio-oil obtained from RWS was lower than that obtained from OPT and hydrocarbon content of bio-oil. Pyrolysis of RWS provided bio-oil with
OPT:RWS. This may be because the bulk density of RWS was higher than contents of carbon and hydrogen higher than the bio-oil obtained from
of OPT and OPT:RWS, consequently less solid particles were carried out OPT. As can be observed, co-pyrolysis of OPT and RWS improved the
from reaction chamber during pyrolyzing. At a higher pyrolysis tem­ hydrocarbon content of bio-oil as compared to OPT only. The carbon
perature, the solids content in bio-oil was also high. This is due to the content of bio-oil was increased at higher pyrolysis temperature, but
fact the biomass particles were more pyrolyzed at higher temperature, hydrogen content tended to decrease. This is due to the decarboxylation
leading to lower weight of that biomass particles and then the biochar and dehydration reactions [62]. In addition, Pattiya and Suttibak [27]
particle can be carried out from reaction chamber. The result of solids reported that the elemental compositions of bio-oil might be influenced
content in bio-oil was similar to the pyrolysis of wood, rice straw, cas­ by the solids content and water content. The high water content in
sava and oil palm biomass [10,55]. The solids present in bio-oil can bio-oil led to high hydrogen content. The bio-oil obtained from this
cause erosion and blockages in pumping or supplying equipment [61]. study consisted mainly of water and oxygen-containing compounds. The
The ash content of bio-oil ranged from 0.3 to 0.44 wt.%. It was found oxygen present in bio-oil is from the cellulose, hemicellulose and lignin
that an increased pyrolysis temperature led to higher ash content in fractions of the biomass [10,27]. The obtained hydrocarbon from this
bio-oil, which was consistent with the results of solids content in bio-oil. study was similar with the result of Alvarez et al. [63]. They found that
This ash was the solid remaining after burning bio-oil under oxidation the hydrocarbon of bio-oil was only 15-27 wt.%. In the case of fast py­
condition. The composition of ash was inorganic components from rolysis, the hydrocarbon content of bio-oil with low water content can
biomass sample. High ash content in bio-oil can cause high wear in range within 60-80 wt.% [52]. The higher hydrocarbon content and
pumps and injectors, as well as deposition and corrosion in equipment lower oxygen content of bio-oil enhances its fuel value. For the content
due to the presence of alkali metals in ash. The density of bio-oil was in of nitrogen and sulphur, it was found that the bio-oil contained low
the range of 1002-1024 kg/m3. The previous studies reported that the amounts of nitrogen and sulphur since the biomass used in this study
density of bio-oil was in the range of 900-1,300 kg/m3, depending on the had low nitrogen and sulphur contents. When the organic elements of
its composition [10,61]. water-free bio-oil samples were presented in the form of atomic ratios in
a Van Krevelen diagram as shown in Fig. 7, it is seen that the atomic
3.3.3. Ultimate analysis results and energy content of bio-oil ratios of H/C and O/C of water-free bio-oil samples were in the ranges of
Table 4 summarizes the ultimate analysis results and higher heating 0.62-1.2 and 0.3-0.9, respectively. The bio-oil samples had low H/C and
values (HHV) of bio-oil. The results reveal that the contents of C, H, N, S high O/C atomic ratios as compared to conventional oils [29,30]. This is
and O of bio-oil were in the ranges 13.8-22.62, 9.12-9.92, 0.07-0.7, because the bio-oil samples contained with low hydrogen and carbon
<0.01 and 68.07-76.21 wt. %, respectively. These results indicated contents, while their oxygen content were relatively high. The HHV of
that biomass types and pyrolysis temperatures had small effect on the bio-oil shown in this study was the HHV of water-free bio-oil as shown in

9
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 5
Main chemical compounds in bio-oil from pyrolysis at 500 ◦ C
Peak area (%)
Compounds
OPT RWS OPT: RWS

Acids
Acetic acid 41.17 37.26 41.19
Propanoic acid 1.87 1.63 1.97
Phenols
Phenol 11.48 3.97 7.5
Phenol, 2-methoxy- 0.95 2.38 1.59
Phenol, 2,6-dimethoxy- 4.14 5.28 5.05
5-tert-Butylpyrogallol 0.68 1.22 1.09
Creosol N/D 1.33 0.76
3,5-Dimethoxy-4-hydroxytoluene 0.8 2.29 1.6
Ketones
2-Propanone, 1-hydroxy- 9.54 8.47 9.19
2-Cyclopenten-1-one 1.76 1.34 1.69
1-Hydroxy-2-butanone 1.29 1.62 1.55
2-Cyclopenten-1-one, 3-methyl- 1.07 0.64 1.08
2-Cyclopenten-1-one, 2,3-dimethyl- 0.83 0.72 1.13
3-Ethyl-2-hydroxy-2-cyclopenten-1-one 0.82 0.69 0.83
2-Cyclopenten-1-one, 2-hydroxy-3-methyl- 2.93 3.34 3.38 Fig. 8. Particle size distributions of biochars from pyrolysis at 500 ◦ C.
Syringylacetone 1.27 1.44 1.37
Aldehydes
2-furan-carboxaldehyde 2.42 2.11 1.63 Table 6
Carbohydrates Proximate analysis results of biochars
Anhydro-sugar 2.85 4.11 3.73
1,4:3,6-Dianhydro-alpha-d-glucopyranose 2.16 1.04 1.61 Proximate analysis (wt.%, as received basis)
Pyrolysis
1,6-anhydro-beta-D-Glucopyranose 4.04 2.79 2.53 Biochar Moisture Volatile Fixed Ash
temperature (◦ C)
Others 7.93 16.33 9.53 content matter carbon content
Total 100 100 100
19.26 ± 68.85 ± 8.27 ±
*N/D = Not detected 400 3.62 ± 0.08
0.55 0.54 0.04
18.29 ± 69.47 ± 8.79 ±
OPT 450 3.45 ± 0.17
0.03 0.08 0.12
Table 4. It was found that the HHV of bio-oil was in the range of
14.15 ± 74.26 ± 8.32 ±
16.1-20.36 MJ/kg3, which was consistent with the hydrocarbon content 500 3.27 ± 0.09
0.3 0.29 0.12
of bio-oil. The HHV of bio-oil obtained from RWS was higher than that of 25.79 ± 68.72 ± 5.48 ±
400 3.45 ± 0.11
OPT and OPT:RWS. The obtained HHV of bio-oil from this study was 0.22 0.13 0.02
also compared to previous studies, including Palamanit et al. [10], Sir­ 20.5 ± 69.82 ± 6.4 ±
RWS 450 3.28 ± 0.16
0.15 0.17 0.15
ijanusorn et al. [26], Pattiya et al. [27], Kabir et al. [52], Chang et al. 19.52 ± 73.4 ± 7.76 ±
[64], and Varma et al. [65]. These studies reported that the HHV of 500 2.92 ± 0.1
0.35 0.29 0.21
bio-oil was in the range 17-39 MJ/kg. The HHV of bio-oil obtained from 24.57 ± 65.55 ± 6.03 ±
400 3.85 ± 0.06
this study was relatively low compared to some previous studies. This is 0.37 0.06 0.06
OPT: 18.16 ± 71.88 ± 6.97 ±
because of the differences in pyrolysis process and biomass composi­ 450 2.98 ± 0.18
RWS 0.36 0.08 0.08
tions. However, the HHV of bio-oil from this study was similar to the 15.17 ± 74.86 ± 7.35 ±
500 2.62 ± 0.12
study of Palamanit et al. [10] who performed slow pyrolysis of oil palm 0.34 0.04 0.04
biomass, and they found that the HHV of bio-oil was in the range of
19-23 MJ/kg. These results make clear that the HHV of bio-oil is lower
than of conventional petroleum fuels (40-50 MJ/kg) [54]. alkali metals with the pyrolysis vapor catalyzed the conversion of the
levoglucosan to light organic compounds such as acetic acid and
3.3.4. Chemical compounds furfural. The amounts of acidic compounds in the obtained bio-oil
Table 5 lists the major chemical compounds of bio-oil obtained from indicated that the pyrolysis product is highly acidic, which is consis­
OPT, RWS and OPT:RWS at pyrolysis temperature of 500 ◦ C. The results tent with the pH of 2.28-2.84, as shown in Table 3. Bio-oil with a low pH
indicated that the main chemical compounds of the bio-oil samples were value is not appropriate for direct use because it encourages aging and
oxygenated compounds such as acetic acid, phenols, 2-Propanone,1-hy­ corrosion [66].
droxy-, and 1,6-anhydro-beta-D-Glucopyranose. The concentrations of
these compounds were in the ranges 37.26-41.17, 3.97-11.48, 8.47-9.54
3.4. Characteristics of biochar
and 2.53-4.04 %, respectively. It is seen that the concentration of each
compound in the bio-oil depended on the biomass type and composition,
3.4.1. Particle size distribution and proximate analysis results
particularly on the lignocellulosic components as depicted in Fig. 3. The
Fig. 8 depicts the particle size distributions of biochar obtained from
obtained results were relatively similar with the previous studies. They
OPT, RWS and OPT:RWS at pyrolysis temperature of 500 ◦ C. It is seen
reported that the chemical compounds in the bio-oil is strongly influ­
that the particle size distribution of most biochars was in the range of
enced by biomass compositions [10,52]. They also indicated that the
0.2-0.3 mm. The results indicate the particle size distribution of biochar
cellulose and hemicellulose in the biomass could decompose to gluco­
was consistent with the particle size distribution of raw biomass sam­
pyranoses (anhydro-sugars) and heterocyclic compounds (furan and
ples. The particle size of biochar was lower due to the shrinkage of
furan derivatives, and acetic acid). In the part of lignin, it was devola­
biomass particle caused by thermal decomposition during pyrolysis. The
tilized to pervade the resulting bio-oil with oxygenated aromatic com­
shrinkage of biochar particle is mainly influenced by the nature of the
pounds such as phenols and phenolic derivatives [52,55]. Besides the
biomass and the pyrolysis conditions [67]. For the proximate analysis
lignin content, the alkali metals of the biomass also influenced the
results of biochar, it is shown Table 6. This table shows the proximate
compounds of the bio-oil. Kabir et al. [52] explained that the contact of
analysis results of the biochar obtained from OPT, RWS, and OPT: RWS

10
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Table 7
Ultimate analysis results and higher heating values of biochars
Ultimate analysis (wt.%, dry basis)
Biochar Pyrolysis temperature (◦ C) HHV (MJ/kg)
Carbon Hydrogen Nitrogen Sulphur *Oxygen

400 74.14 ± 0.16 3.16 ± 0.03 0.58 ± 0.01 0.06 13.48 ± 0.18 27.65 ± 0.01
OPT 450 75.12 ± 0.1 3.01 ± 0.07 0.56 ± 0.01 0.06 12.15 ± 0.22 27.94 ± 0.14
500 77.45 ± 0.24 2.38 ± 0.06 0.51 0.06 11 ± 0.22 28.05 ± 0.11
400 71.63 ± 0.24 3.28 ± 0.05 0.4 0.03 18.98 ± 0.2 26.35 ± 0.07
RWS 450 77.1 ± 0.21 3.11 ± 0.01 0.39 ± 0.01 0.03 ± 0.01 12.75 ± 0.2 28.73 ± 0.08
500 80.38 ± 0.43 2.56 ± 0.01 0.41 ± 0.01 0.04 8.62 ± 0.45 29.6 ± 0.21
400 73.29 ± 0.36 3.53 ± 0.08 0.45 ± 0.01 0.04 16.48 ± 0.41 27.54 ± 0.18
OPT:RWS 450 76.55 ± 0.12 3.01 ± 0.02 0.51 ± 0.01 0.05 12.7 ± 0.11 28.4 ± 0.23
500 78.60 ± 0.09 2.45 ± 0.03 0.46 ± 0.02 0.05 10.89 ± 0.18 28.57 ± 0.04
*
Oxygen by difference.

Table 8
BET surface areas and average pore diameters of biomasses and biochars (500

C)
Type of biomass/ BET surface area (m2/g) Average pore diameters
biochar (nm)

OPT 3.1 11.46


RWS 2.0 13.9
OPT:RWS 2.58 12.68
OPT biochar 2.1 21.2
RWS biochar 1.5 32.8
OPT:RWS biochar 1.3 24.63

biomasses and pyrolysis temperatures. It is seen that the C, H, N, S and O


content of the biochar was in the range of 74.14-80.38, 2.38-3.53, 0.41-
0.58, 0.03-0.06 and 8.62-18.98 wt.%, respectively. The results of ulti­
mate analysis were consistent with the proximate analysis results as
indicated by the relation between fixed carbon content and volatile
matter, and carbon and hydrogen content. The carbon content of OPT
Fig. 9. Van Krevelen diagram (H/C and O/C atomic ratios) of biomass, biochar and OPT: RWS was lower than that of RWS. Pyrolysis temperature
(500 ◦ C) and coal. showed more significant effect on biochar. These results clearly showed
that the pyrolysis of biomass produced biochar with a high carbon
at various pyrolysis temperatures. As can be seen the moisture content, content and a low oxygen content, which is similar to the results from
volatile matter, fixed carbon content, and ash content of the biochar previous studies [10]. As depicted by Van Krevelen diagram in Fig. 9, the
were in the ranges 2.92-3.85, 14.15-25.79, 65.55-74.86 and 5.48-8.79 ratio of H/C and O/C of biochar was clearly improved as compared to
wt.%, respectively. These results indicate that differences in the type raw biomass. The reduction of hydrogen and oxygen content at higher
of oil palm biomass and pyrolysis temperature affected the composition pyrolysis temperature was attributed to the dehydration and decar­
of the biochar. Based on the obtained results, it is observed that the boxylation reactions [55]. That ratios of biochar were close to the values
moisture content of biochar samples was not zero after pyrolysing, of each coal, thus the biochar normally can be called “biocoal”. In
which was similar to previous studies [10,52]. They reported that the practice, the lower ratio of O/C and H/C is favorable for solid fuel
moisture content of biochars was in the range of 2.40-4.42 wt.%. The application via combustion or co-firing. The nitrogen and sulfur content
volatile matter of biochar decreased with pyrolysis temperature. This is of biochar was relatively low. However, as compared to the raw biomass
because the volatile matter was more decomposed to low molecular the biochar had higher nitrogen content. This is due to the combination
weight liquid and gas at higher temperature [10]. However, the volatile effect of nitrogen into complex structures and nitrogen bond is also
matter of the biochar remained between 14.15-25.75 wt.%, indicating resistant to heat and not easily volatilized [68]. The higher carbon
incomplete pyrolysis of the biomass under the investigated pyrolysis content and lower oxygen content of the biochar led to superior HHV
temperatures and time. This is because the biomass samples contained
some lignin, which decomposes at high temperatures. The lower volatile
matter led to higher fixed carbon content of biochar. The high fixed
carbon content of biochar is favorable for solid fuel application due to
high energy content [10,55]. The high fixed carbon content of biochar
also helps improve stability during storage as compared to the raw
biomass, which is easily biologically decomposed during storage. The
ash content of OPT was higher than RWS and OPT: RWS, which was
consistent with the gross component of raw biomass. The ash in biomass
and biochar is a non-volatile matter and non-combustible components.
The increase in ash content of biochar resulted from the destructive
volatilization of lignocellulose components at higher temperature [68].

3.4.2. Ultimate analysis results and energy content of biochar


Table 7 shows the elemental compositions of the biochar at different
Fig. 10. FTIR spectra of biochars from pyrolysis at 500 ◦ C.

11
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

Fig. 11. SEM photographs of (a) OPT, (b) OPT biochar, (c) RWS, (d) RWS biochar, (e) OPT: RWS, (f) OPT: RWS biochar. The biochars are from pyrolysis at 500 ◦ C.

[3]. As can be seen in Table 8, the HHV of biochar was in the range of Thus, the low moisture content, low ratio of O/C and H/C, combined
26.35-29.6 MJ/kg. The HHV of RWS biochar was higher than OPT and with high HHV of biochar are desirable for solid fuel applications for
OPT: RWS biochar, which is consistent with the HHV of biomass. The both heat and power generation since these properties help to reduce the
higher carbon content and lower oxygen content of the biochar led to airborne waste and energy loss [70].
superior HHV as compared to raw biomass. Previous studies showed that
the HHV of biochar obtained from woody and non-woody biomasses was 3.4.3. FTIR analysis results
in the range of 14-33 MJ/kg, depending on the type of used biomass and The FTIR analysis results of the biochar obtained from pyrolysis of
the pyrolysis conditions. The previous studies also indicated that the OPT, RWS and OPT:RWS at temperature of 500 ◦ C are shown in Fig. 10.
HHV of biochar from woody biomass was better than of biochar from This figure showed that there were many components present in the
non-woody samples [10,52,53,64]. As compared to the hydrochar, the obtained results. This is because the biochar still contained structures of
HHV and carbon content of biochar was better than hydrochar [69]. carbon, hydrogen and oxygen, which mostly remained in the form of

12
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

degradation of the lignocellulosic components including cellulose,


hemicelluloses and lignin. At higher temperature, it also facilitates the
release of volatile materials and created more pores, resulting in larger
pore size of biochar [10]. Previous study also reported that at heating
rate higher than 10 ◦ C/min, the volatile vapor was generated inside the
biomass particles and then it released from surface, which resulted in
open fiber structure leading to the formation of cavities. The high porous
structure of biochar may be beneficial for bio-filter applications and soil
mixing, in addition to use as solid fuel [8,73].

3.5. Pyrolysis gas composition and heating value

Fig. 12 shows the main compositions and heating values of the py­
rolysis gas products obtained from pyrolysis temperature of 500 ◦ C. The
results showed that the pyrolysis gas mainly contained H2, CO, CH4 and
CO2. The concentrations of these gases were in the ranges 4.42-9, 1.34-
17.96, 4.76-7.62, and 6.8-29.54 vol. %, respectively. With obtained
concentration of combustible pyrolysis gases, the heating value of py­
Fig. 12. Composition and heating value of pyrolysis gas at 500 ◦ C.
rolysis gas was in the range of 3.23-6.7 MJ/m3. The obtained pyrolysis
gas contained lower concentrations of combustible gases as compared to
lignin. The result of FTIR are consistent with the proximate and ultimate the results from previous studies [53,61,63]. The concentration of
analysis results of biochar discussed previously. The spectra of each combustible pyrolysis gas was relatively low due to the slow pyrolysis
biochar sample were not different because of the elemental composition process and batch runs used in this study. In addition, the N2 gas flow
of biochar samples was relatively similar as indicated by the content of rate supplied to the reactor was also relatively high. Considering the
carbon, hydrogen and oxygen. The peaks can be explained as follows. effects of biomass type on pyrolysis gas composition, the results showed
The first peak appeared at 3385-3386 cm− 1 and was attributed to the that pyrolysis of OPT provided more H2, CO and CH4 than that obtained
stretching of OH groups [69]. It is also attributed to acceleration in from RWS and OPT:RWS. This was attributed to the lignocellulosic
dehydration reaction of biomass. The small peaks at 2850-3011 cm-1 compositions of biomass samples [10,61]. The amount of hemicellulose
were associated with the C-H stretching vibration of aliphatic and aro­ combined with cellulose in OPT was lower than in RWS. Consequently,
matics structures. The observed peak at 1685 cm− 1 is attributed the RWS is mostly converted as condensable vapor, while the incondensable
presence of carbonyl and carboxyl groups in carbohydrates [71]. The vapor of OPT was high. Thus, at pyrolysis temperature of 500 ◦ C the OPT
aromatic C = C ring stretching vibration occurred at 1565–1566 cm-1 provided higher concentration of H2, CO and CO2 than RWS and OPT:
[55]. At 1378-1402 cm-1, the peak is assigned mainly to stretching vi­ RWS. This result was also consistent with the TGA results. The result of
brations of aliphatic C–H and CH2 bending in biochar, respectively. The this study are similar to Paenpong and Pattiya [61], who showed that
bands in the range 1260-1261 cm-1 represent the stretching of aromatic pyrolysis of biomass with high hemicellulose and cellulose led to high
C–O and phenolic OH. The weak vibrations of C-H bond in aromatic and concentration of combustible gases. Moreover, gas composition may be
heteroaromatic compounds are visible as a band between 611-873 cm− 1 influenced by the secondary reactions of condensable vapor with ash
[71,72]. during the process [73,74]. This result led to high gas yield and gas
concentration such in the case of OPT. However, it is noted that the
3.4.4. Surface morphology and surface area composition and concentration of the pyrolysis gas also depend on other
Fig. 11(a)–(f) show the surface features of the biomass and biochar factors such as the reactor type, pyrolysis temperature, pyrolysis type
obtained from OPT, RWS, and OPT:RWS at a pyrolysis temperature of and N2 gas flow rate [10,74].
500 ◦ C. These images clearly showed the difference between biomass
and biochar surface features. The surface of biomass samples was rela­ 4. Conclusion
tively rough with small pores as shown in Fig. (a), (c) and (e). It is seen
that the surface of RWS biomass was relatively rougher than OPT The yields and characteristics of bio-oil, biochar and pyrolysis gas
biomass. After pyrolysis process, the morphology of biochar revealed obtained from pyrolysis and co-pyrolysis of RWS and OPT were inves­
pores created over the surface, with the size and shape of the pores tigated by using an agitated pyrolysis reactor. The co-pyrolysis was
clearly seen in Fig. (b), (d) and (f). The RWS biochar had rough surface performed at the mixing ratio of 50:50 (wt.). The biomass samples were
with non-uniform shape of pores as compared to OPT biochar. The shape pyrolyzed at temperatures of 400, 450, and 500 ◦ C under specific con­
of the OPT biochar pores at the surface appears like a honeycomb with ditions. The yield and characteristics of bio-oil, biochar and pyrolysis
cylindrical and polygonal pores. The average pore diameter of RWS gas were then determined and analyzed. The results indicated the py­
biochar was bigger than OPT biochar. The surface feature of OPT bio­ rolysis temperatures had more effect on the yield and characteristics of
char obtained from this study was similar to the result of Palamanit et al. bio-oil and biochar than biomass types. The yields of the bio-oil, biochar
[10], and Bensidhom et al. [53]. The creation of large pores is caused by and pyrolysis gas were in the ranges of 38.5-46.5, 22.27-28.68 and 30.1-
the volatilization of organic compounds or lignocellulosic components. 36.9 wt.%, respectively. Pyrolysis of RWS at 500 ◦ C provided the highest
When the pore was larger, the number of pores might be lower, leading yield. The bio-oil had high water content and low pH. The main com­
to reduced surface area [55]. The BET surface area of biomass and pounds of bio-oil were oxygenated compounds such as acetic acid,
biochar was determined and shown in Table 8. It is observed that the phenols and 2-Propanone,1-hydroxy- as indicated by GC-MS. The high
surface area of OPT, RWS and OPT:RWS biochar was lower as compared water content of bio-oil led to low HHV (16.1-20.36 MJ/kg). The bio­
to the raw biomass sample. This is because the thermal decomposition of char had high carbon and low oxygen contents, matching proximate and
biomass led to larger pore size, in which the number of pores might be ultimate analysis and FTIR. The HHV of biochar was in the range of
lower. Comparing between biomass types, it was found that the BET 26.35-29.6 MJ/kg. Besides this, the biochar also had porous structure,
surface area of OPT was higher than RWS and OPT:RWS. This is prob­ which was indicated by SEM and BET. The pyrolysis gas obtained from
ably due to thermal decomposition of RWS was better than OPT as 500 ◦ C mainly contained CO2, CO, H2 and CH4, and its heating value was
shown in TGA graph. The large pores are caused by the progressive in the range 3.23-6.7 MJ/m3. Based on these results, OPT and RWS are

13
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

alternative biomasses for biofuel, biochemical and bio-product pro­ [15] B. Biswas, N. Pandey, Y. Bisht, R. Singh, J. Kumar, T. Bhaskar, Pyrolysis of
agricultural biomass residues: Comparative study of corn cob, wheat straw, rice
duction via pyrolysis and co-pyrolysis processes.
straw and rice husk, Bioresour. Technol. 237 (2017) 57–63, https://doi.org/
10.1016/j.biortech.2017.02.046.
CRediT authorship contribution statement [16] M.A. Sukiran, S.K. Loh, N.A. Bakar, Production of bio-oil from fast pyrolysis of oil
palm biomass using fluidised bed reactor, J. Energy Resour. Technol. 6 (2016)
1–11.
Patipan Sakulkit: Conceptualization, Investigation, Visualization, [17] A.A. Salema, R.M.W. Ting, Y.K. Shang, Pyrolysis of blend (oil palm biomass and
Writing - original draft. Arkom Palamanit: Conceptualization, Re­ sawdust) biomass using TG-MS, Bioresour. Technol. (2018).
[18] F.N. Ani, R. Zailani, Characteristics of pyrolysis oil and char from oil palm shells,
sources, Writing - review & editing, Supervision. Racha Dejchanchai­
Dev. Thermochem. Biomass convers, Springer Netherlands, Dordrecht, 1997,
wong: Writing - review & editing, Supervision. Prasert pp. 425–432.
Reubroycharoen: Writing - review & editing, Supervision. [19] P. Basu, Biomass characteristics, Biomass Gasification Design Handbook, 2010.
[20] C. Sheng, J.L.T. Azevedo, Modeling biomass devolatilization using the chemical
percolation devolatilization model for the main components, Proc. Combust. Inst.
Declaration of Competing Interest 29 (2002) 407–414, https://doi.org/10.1016/S1540-7489(02)80054-2.
[21] A. Demirbas, Relationships between lignin contents and heating values of biomass,
Energy Convers. Manage. 42 (2001) 183–188, https://doi.org/10.1016/S0196-
The authors declare that they have no known competing financial 8904(00)00050-9.
interests or personal relationships that could have appeared to influence [22] H.K. Goering, P.J. Van Soest, Forage fiber analysis, apparatus, reagents,
procedures, and some applications. Agriculture handbook, ARS-USDA,
the work reported in this paper Washington, DC, 1970.
[23] P.J. Van Soest, J.B. Robertson, B.A. Lewis, Methods for dietary fiber, neutral
Acknowledgements detergent fiber, and nonstarch polysaccharides in relation to animal nutrition,
J. Dairy. Sci. 74 (1991) 3583–3597, https://doi.org/10.3168/jds.S0022-0302(91)
78551-2.
The authors would like to thank to the Interdisciplinary Graduate [24] M.T. Reza, M.H. Uddin, J.G. Lynam, S.K. Hoekman, C.J. Coronella, Hydrothermal
School of Energy Systems, Prince of Songkla University (Contract carbonization of loblolly pine: reaction chemistry and water balance, Biomass
Convers. Biorefin. 4 (2014) 311–321, https://doi.org/10.1007/s13399-014-0115-
number: 1-2018/06) and Provincial Electricity Authority (PEA) of 9.
Thailand for their financial support. This study was also partially sup­ [25] H.S. Kambo, A. Dutta, Comparative evaluation of torrefaction and hydrothermal
port by Graduate Student Fund (Contract number: 033/2562) from carbonization of lignocellulosic biomass for the production of solid biofuel, Energy
Convers. Manag. 105 (2015) 746–755, https://doi.org/10.1016/j.
Energy Conservation Promotion Fund, Energy Policy and Planning Of­ enconman.2015.08.031.
fice (EPPO), Ministry of Energy, Thailand. The authors thank to [26] S. Sirijanusorn, K. Sriprateep, A. Pattiya, Pyrolysis of cassava rhizome in a counter-
Research and Development Office (RDO), Prince of Songkla University, rotating twin screw reactor unit, Bioresour. Technol. 139 (2013) 343–348, https://
doi.org/10.1016/j.biortech.2013.04.024.
and Assoc. Prof. Dr. Seppo Karrila for support an English correction.
[27] A. Pattiya, S. Suttibak, Influence of a glass wool hot vapour filter on yields and
properties of bio-oil derived from rapid pyrolysis of paddy residues, Bioresour.
References Technol. 116 (2012) 107–113, https://doi.org/10.1016/j.biortech.2012.03.116.
[28] A. Pattiya, S. Suttibak, Production of bio-oil via fast pyrolysis of agricultural
residues from cassava plantations in a fluidised-bed reactor with a hot vapour
[1] L. Plante, N.P. Sheehan, P. Bier, K. Murray, K. Quell, C. Ouellette, E. Martinez,
filtration unit, J. Anal. Appl. Pyrol. 95 (2012) 227–235, https://doi.org/10.1016/j.
Bioenergy from biofuel residues and waste, Water Environ. Res. (2019), https://
jaap.2012.02.010.
doi.org/10.1002/wer.1214.
[29] S. Jamilatun, Rochmadi Budhijanto, A. Yuliestyan, H. Hadiyantod, A. Budimanb,
[2] A. Ahmed, M.S. Abu Bakar, R. Hamdani, Y.K. Park, S.S. Lam, R.S. Sukri, M. Aslam,
Comparative analysis between pyrolysis products of spirulina platensis biomass
Valorization of underutilized waste biomass from invasive species to produce
and its residues, Int. J. Renewable Energy Dev. 8 (2019) 133–140, https://doi.org/
biochar for energy and other value-added applications, Environ. Res. 186 (2020),
10.14710/ijred.8.2.133-140.
109596, https://doi.org/10.1016/j.envres.2020.109596.
[30] B.B. Uzun, E.A. Varol, E. Pütün, Pyrolysis: A sustainable way from biomass to
[3] P. Shrivastava, P. Khongphakdi, A. Palamanit, A. Kumar, P. Tekasakul,
biofuels and biochar (n.d.), Biochar (2020) 239–265, https://doi.org/10.1017/
Investigation of physicochemical properties of oil palm biomass for evaluating
9781316337974.013.
potential of biofuels production via pyrolysis processes, Biomass Convers. Biorefin.
[31] A. Ahmed, M.S. Abu Bakar, A.K. Azad, R.S. Sukri, N. Phusunti, Intermediate
(2020), https://doi.org/10.1007/s13399-019-00596-x.
pyrolysis of Acacia cincinnata and Acacia holosericea species for bio-oil and
[4] R.K. Mishra, K. Mohanty, Pyrolysis kinetics and thermal behavior of waste sawdust
biochar production, Energy Convers. Manage. 176 (2018) 393–408.
biomass using thermogravimetric analysis, Bioresour. Technol. 251 (2018) 63–74,
[32] S. Nanda, J. Mohammad, S.N. Reddy, J.A. Kozinski, A.K. Dalai, Pathways of
https://doi.org/10.1016/j.biortech.2017.12.029.
lignocellulosic biomass conversion to renewable fuels, Biomass Convers. Biorefin. 4
[5] Rubber Authority of Thailand, Academic information of rubber 2018, 2018
(2014) 157–191, https://doi.org/10.1007/s13399-013-0097-z.
(Accessed 22 June 2019, https://km.raot.co.th/book/read-product/230.
[33] N. Abdoulmoumine, A. Kulkarni, S. Adhikari, Effects of temperature and
[6] Office of Agricultural Economics: Agricultural Statistics of Thailand 2018, Office of
equivalence ratio on mass balance and energy analysis in loblolly pine oxygen
Agricultural Economics, Ministry of Agriculture and Cooperatives of Thailand,
gasification, Energy Sci. Eng. 4 (2016) 256–268, https://doi.org/10.1002/
2018 (Accessed 22 June 2019), http://www.oae.go.th/assets/portals/1/ebookcate
ese3.124.
gory/27_yearbook2561/.
[34] A. Odeha, S. Ogbeideb, C. Okieimenc, Elucidation of the influence of coal
[7] S. Kaewluan, S. Pipatmanomai, Gasification of high moisture rubber woodchip
properties on coal-char reactivity: A Look at Southern Hemisphere Coals, J. Chem.
with rubber waste in a bubbling fluidized bed, Fuel Process. Technol. 92 (2011)
Eng. Process Technol. 9 (2018) 2, https://doi.org/10.4172/2157-7048.1000380.
671–677, https://doi.org/10.1016/j.fuproc.2010.11.026.
[35] C. Mamimin, P. Thongdumyu, A. Hniman, P. Prasertsan, T. Imai, S. O-Thong,
[8] A. Shariff, R. Hakim, N. Abdullah, Rubber wood as potential biomass feedstock for
Simultaneous thermophilic hydrogen production and phenol removal from palm
biochar via slow pyrolysis, Int. J. Chem. Biol. Eng. 10 (2016) 12.
oil mill effluent by Thermoanaerobacterium-rich sludge, Int. J. Hydrogen Energy.
[9] T.M.I. Mahlia, N. Ismail, N. Hossain, A.S. Silitonga, A.H. Shamsuddin, Palm oil and
37 (2012) 15598–15606, https://doi.org/10.1016/j.ijhydene.2012.04.062.
its wastes as bioenergy sources: a comprehensive review, Environ. Sci. Pollut. Res.
[36] P. Suwannakuta, A study on biomass gasification in spout-fluid bed. Master Thesis
(2019), https://doi.org/10.1007/s11356-019-04563-x.
No. ET-02-22, School of environment resources and development, Asian institute of
[10] A. Palamanit, P. Khongphakdi, Y. Tirawanichakul, N. Phusunti, Investigation of
technology, Bangkok, Thailand, 2002.
yields and qualities of pyrolysis products obtained from oil palm biomass using an
[37] P. Madhu, T. Livingston, H. Kanagasabapathy, Flash pyrolysis of lemon grass
agitated bed pyrolysis reactor, Biofuel Res. J. 24 (2019) 1065–1079, https://doi.
(Cymbopogon flexuosus) for bio-oil production in an electrically heated fluidized
org/10.18331/BRJ2019.6.4.3.
bed reactor, Waste Biomass Valoriz. 9 (2017) 1037–1046, https://doi.org/
[11] P. Khongphakdi, A. Palamanit, N. Phusunti, Y. Tirawanichakul, P. Shrivastava,
10.1007/s12649-017-9872-6.
Evaluation of oil palm biomass potential for bio-oil production via pyrolysis
[38] E.B. Jose, T. Bhaskar, Biomass and Biofuels: Advanced biorefineries for sustainable
processes, Int. J. Integr. Care. 11 (2019) 45–52.
production and distribution, CRC Press, 2015, https://doi.org/10.1201/b18398.
[12] A. Tursi, A review on biomass: importance, chemistry, classification, and
[39] A. Pattiya, Bio-oil production via fast pyrolysis of biomass residues from cassava
conversion, Biofuel Res. J. 22 (2019) 962–979, https://doi.org/10.18331/
plants in a fluidised-bed reactor, Bioresour. Technol. 102 (2011) 1959–1967,
BRJ2019.6.2.3.
https://doi.org/10.1016/j.biortech.2010.08.117.
[13] A.I. Ferreiro, P. Giudicianni, C.M. Grottola, M. Rabaçal, M. Costa, R. Ragucci,
[40] M. Asadullah, M.A. Rahman, M.M. Ali, M.A. Motin, M.B. Sultan, M.R. Alam, M.
Unresolved issues on the kinetic modeling of pyrolysis of woody and nonwoody
S. Rahman, Jute stick pyrolysis for bio-oil production in fluidized bed reactor,
biomass fuels, Energy Fuels. 31 (2017) 4035–4044, https://doi.org/10.1021/acs.
Bioresour. Technol. 99 (2008) 44–50, https://doi.org/10.1016/j.
energyfuels.6b03445.
biortech.2006.12.002.
[14] E.R. Widjaya, G. Chen, L. Bowtell, C. Hills, Gasification of non-woody biomass: A
[41] R. Sakthivel, K. Ramesh, P.M. Shameer, R. Purnachandran, A complete analytical
literature review, Renewable Sustainable Energy Rev. 89 (2018) 184–193, https://
characterization of products obtained from pyrolysis of wood barks of calophyllum
doi.org/10.1016/j.rser.2018.03.023.

14
P. Sakulkit et al. Journal of Environmental Chemical Engineering 8 (2020) 104561

inophyllum, Waste Biomass Valoriz. (2018), https://doi.org/10.1007/s12649-018- [59] V.K. Guda, H. Toghiani, Altering bio-oil composition by catalytic treatment of pine
0236-7. wood pyrolysis vapors over zeolites using an auger - packed bed integrated reactor
[42] M. Asadullah, M.A. Rahman, M.M. Ali, M.S. Rahman, M.A. Motin, M.B. Sultan, M. system, Biofuel Res. J. 11 (2016) 448–457, https://doi.org/10.18331/
R. Alam, Production of bio-oil from fixed bed pyrolysis of bagasse, Fuel 86 (2007) BRJ2016.3.3.4.
2514–2520, https://doi.org/10.1016/j.fuel.2007.02.007. [60] T. Chorazy, J. Čáslavský, V. Žvaková, J. Raček, P. Hlavínek, Characteristics of
[43] P. Pimenidou, V. Dupont, Characterisation of palm empty fruit bunch (PEFB) and pyrolysis oil as renewable source of chemical materials and alternative fuel from
pinewood bio-oils and kinetics of their thermal degradation, Bioresour. Technol. the sewage sludge treatment, Waste Biomass Valoriz. (2019), https://doi.org/
109 (2012) 198–205, https://doi.org/10.1016/j.biortech.2012.01.020. 10.1007/s12649-019-00735-5.
[44] S.D. Stefanidis, K.G. Kalogiannis, E.F. Iliopoulou, C.M. Michailof, P.A. Pilavachi, A. [61] C. Paenpong, A. Pattiya, Effect of pyrolysis and moving-bed granular filter
A. Lappas, A study of lignocellulosic biomass pyrolysis via the pyrolysis of temperatures on the yield and properties of bio-oil from fast pyrolysis of biomass,
cellulose, hemicellulose and lignin, J. Anal. Appl. Pyrol. 105 (2014) 143–150, J. Anal. Appl. Pyrol. 119 (2016) 40–51, https://doi.org/10.1016/j.
https://doi.org/10.1016/j.jaap.2013.10.013. jaap.2016.03.019.
[45] X. Qu, P. Liang, Z. Wang, R. Zhang, D. Sun, X. Gong, Pilot development of [62] J. Akhtar, N.S. Amin, A review on operating parameters for optimum liquid oil
polygeneration process of circulating fluidized bed combustion combined with coal yield in biomass pyrolysis, Renew. Sustain. Energy. Rev. 16 (5) (2012) 101–109,
pyrolysis, Chem. Eng. Technol. 34 (2011) 61–68, https://doi.org/10.1002/ https://doi.org/10.1016/j.rser.2012.05.033.
ceat.201000202. [63] J. Alvarez, M. Amutio, G. Lopez, J. Bilbao, M. Olazar, Fast co-pyrolysis of sewage
[46] S.S. Kim, H.V. Ly, J. Kim, J.H. Choi, H.C. Woo, Thermogravimetric characteristics sludge and lignocellulosic biomass in a conical spouted bed reactor, Fuel 159
and pyrolysis kinetics of Alga Sagarssum sp. Biomass, Bioresour. Technol. 139 (2015) 810–818, https://doi.org/10.1016/j.fuel.2015.07.039.
(2013) 242–248, https://doi.org/10.1016/j.biortech.2013.03.192. [64] G. Chang, Y. Huang, J. Xie, H. Yang, H. Liu, X. Yin, C. Wu, The lignin pyrolysis
[47] C. Quan, N. Gao, Q. Song, Pyrolysis of biomass components in a TGA and a fixed- composition and pyrolysis products of palm kernel shell, wheat straw, and pine
bed reactor: thermochemical behaviors, kinetics, and product characterization, sawdust, Energy Convers. Manage. 124 (2016) 587–597, https://doi.org/10.1016/
J. Anal. Appl. Pyrol. 121 (2016) 84–92, https://doi.org/10.1016/j. j.enconman.2016.07.038.
jaap.2016.07.005. [65] A.K. Varma, L.S. Thakur, R. Shankar, P. Mondal, Pyrolysis of wood sawdust: Effects
[48] T. Wang, Y. Chen, J. Li, Y. Xue, J. Liu, M. Mei, S. Chen, Co-pyrolysis behavior of of process parameters on products yield and characterization of products, Waste
sewage sludge and rice husk by TG-MS and residue analysis, J. Cleaner Prod. 250 Management. 89 (2019) 224–235, https://doi.org/10.1016/j.
(2019), 119557, https://doi.org/10.1016/j.jclepro.2019.119557. wasman.2019.04.016.
[49] A. Demirbaş, Mechanisms of liquefaction and pyrolysis reactions of biomass, [66] G. Kabir, B.H. Hameed, Recent progress on catalytic pyrolysis of lignocellulosic
Energy Convers. Manage. 41 (2000) 633–646, https://doi.org/10.1016/S0196- biomass to high- grade bio-oil and bio-chemicals, Renew. Sustain. Energy. Rev. 70
8904(99)00130-2. (2016) 945–967, https://doi.org/10.1016/j.rser.2016.12.001.
[50] T. Kan, V. Strezov, T.J. Evans, Lignocellulosic biomass pyrolysis: A review of [67] E. Cetin, B. Moghtaderi, R. Gupta, T. Wall, Influence of pyrolysis conditions on the
product properties and effects of pyrolysis parameters, Renew. Sustain. Energy. structure and gasification reactivity of biomass chars, Fuel 83 (2004) 2139–2150,
Rev. 57 (2016) 1126–1140, https://doi.org/10.1016/j.rser.2015.12.185. https://doi.org/10.1016/j.fuel.2004.05.008.
[51] R. Fahmi, A.V. Bridgwater, I. Donnison, N. Yates, J.M. Jones, The effect of lignin [68] R. Calvelo Pereira, J. Kaal, M. Camps Arbestain, R. Pardo Lorenzo, W. Aitkenhead,
and inorganic species in biomass on pyrolysis oil yields, quality and stability, Fuel M. Hedley, J.A. Maciá-Agulló, Contribution to characterisation of biochar to
87 (2008) 1230–1240, https://doi.org/10.1016/j.fuel.2007.07.026. estimate the labile fraction of carbon, Organic Geochem. 42 (2011) 1331–1342,
[52] G. Kabir, A.T.M. Din, B.H. Hameed, Pyrolysis of oil palm mesocarp fiber and palm https://doi.org/10.1016/j.orggeochem.2011.09.002.
frond in a slow-heating fixed-bed reactor: a comparative study, Bioresour. Technol. [69] M. Liang, K. Zhang, P. Lei, B. Wang, C.M. Shu, B. Li, Fuel properties and
241 (2017) 563–572, https://doi.org/10.1016/j.biortech.2017.05.180. combustion kinetics of hydrochar derived from co-hydrothermal carbonization of
[53] G. Bensidhom, A.B. Hassen-Trabelsia, K. Alper, M. Sghairoun, K. Zaafouri, tobacco residues and graphene oxide, Biomass Convers. Biorefin. (2019), https://
I. Trabelsi, Pyrolysis of Date palm waste in a fixed-bed reactor: Characterization of doi.org/10.1007/s13399-019-00408-2.
pyrolytic products, Bioresour. Technol. 247 (2018) 363–369, https://doi.org/ [70] Z. Liu, A. Quek, S. Kent Hoekman, R. Balasubramanian, Production of solid biochar
10.1016/j.biortech.2017.09.066. fuel from waste biomass by hydrothermal carbonization, Fuel 103 (2013) 943–949,
[54] S. Şensöz, D. Angın, S. Yorgun, O.M. Koçkar, Bio-oil production from an oilseed https://doi.org/10.1016/j.fuel.2012.07.069.
crop: fixed-bed pyrolysis of rapeseed (Brassica napus L.), Energy Sources. 22 [71] M.Z. Bavariani, A. Ronaghi, R. Ghasemi, Influence of pyrolysis temperatures on
(2000) 891–899, https://doi.org/10.1080/00908310051128255. FTIR analysis, nutrient bioavailability, and agricultural use of poultry manure
[55] N. Bhattacharjee, A.B. Biswas, Pyrolysis of Alternanthera philoxeroides (alligator biochars, Commun. Soil Sci. Plant Anal. (2019) 1–10, https://doi.org/10.1080/
weed): Effect of pyrolysis parameter on product yield and characterization of liquid 00103624.2018.1563101.
product and bio char, J. Energy Inst. 91 (2018) 605–618, https://doi.org/10.1016/ [72] Y.A. Elnour, A.A. Alghyamah, H.M. Shaikh, A.M. Poulose, S.M. Al-Zahrani, A. Anis,
j.joei.2017.02.011. M.I. Al-Wabel, Effect of pyrolysis temperature on biochar microstructural
[56] H. Vasu, C.F. Wong, N.R. Vijiaretnam, Y.Y. Chong, Thangalazhy Gopakumar, evolution, physicochemical characteristics, and its influence on biochar/
S. Gan, L.Y. Lee, H.K. Ng, Insight into co pyrolysis of palm kernel shell (PKS) with polypropylene composites, Appl. Sci. 9 (2019) 1149, https://doi.org/10.3390/
palm oil sludge (POS): Effect on bio oil yield and properties, Waste Biomass app9061149.
Valoriz. (2019), https://doi.org/10.1007/s12649-019-00852-1. [73] F. Abnisa, W.M.A. Wan Daud, A review on co-pyrolysis of biomass: An optional
[57] T. Chen, C. Wu, R. Liu, W. Fei, S. Liu, Effect of hot vapor filtration on the technique to obtain a high-grade pyrolysis oil, Energy Convers. Manage. 87 (2014)
characterization of bio-oil from rice husks with fast pyrolysis in a fluidized-bed 71–85, https://doi.org/10.1016/j.enconman.2014.07.007.
reactor, Bioresour. Technol. 102 (2011) 6178–6185, https://doi.org/10.1016/j. [74] M. Tayyab, A. Noman, W. Islam, S. Waheed, Y. Arafat, F. Ali, M. Zaynab, S. Lin,
biortech.2011.02.023. H. Zhang, W. Lin, Bioethanol production from lignocellulosic biomass by
[58] N. Gómez, S.W. Banks, D.J. Nowakowski, J.G. Rosas, J. Cara, M.E. Sánchez, A. environment-friendly pretreatment methods: A review, Appl. Ecol. Env. Res. 16
V. Bridgwater, Effect of temperature on product performance of a high ash biomass (2018) 225–249, https://doi.org/10.15666/aeer/1601_225249.
during fast pyrolysis and its bio-oil storage evaluation, Fuel Process. Technol. 172
(2018) 97–105, https://doi.org/10.1016/j.fuproc.2017.11.021.

15

You might also like