You are on page 1of 9

Renewable Energy 143 (2019) 112e120

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Gasification of waste biomass for hydrogen production: Effects of


pyrolysis parameters
Phuet Prasertcharoensuk, Steve J. Bull, Anh N. Phan*
School of Engineering, Newcastle University, Newcastle Upon Tyne, UK

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

Article history: Understanding the behaviour of pyrolysis is crucial in gasification of high volatile content lignocellulosic
Received 6 December 2018 material. In this study, parameters affecting the properties of volatiles and char that are feedstock for the
Received in revised form gasification step were studied to determine and optimize operating conditions in pyrolysis for high
21 March 2019
quality syngas/hydrogen production. A uniform temperature profile was obtained at particle sizes up to
Accepted 3 May 2019
0.5 cm3. Pyrolysis temperatures in the range 600e900  C significantly influence the char properties, i.e.
Available online 4 May 2019
increasing surface area and total pore size up to 2.5e3 times when increasing temperature, which in turn
enhances the gas-solid reactions occurring in the gasification process. Increasing pyrolysis temperatures,
Keywords:
Waste wood
i.e. above 700  C fully decomposed unstable compounds, i.e. levoglucosan and their derivatives, but
Pyrolysis promoted the formation of phenolic compounds. Around 41% reduction in surface area and total pore
Gasification volume of the char was observed when increasing particle size from 0.5 to 2 cm3. At pyrolysis temper-
Hydrogen atures above 800  C and particle size of 0.5e1 cm3 with a controlled amount of steam, i.e. 5.7 steam to
Particle size carbon in biomass (S/C) molar ratio, the H2 content in the gas phase increased to 67 mol% from 49 mol%
Temperature and the aromatic compounds in the gas stream decreased up to 50.6%.
Crown Copyright © 2019 Published by Elsevier Ltd. All rights reserved.

1. Introduction temperature range (250e400  C) [1,2] potentially leading to the


formation of high molecular weight compounds (known as tar)
Among available renewable resources, only biomass contains which are difficult to break down in later stages of processing and
the carbon and hydrogen that can be converted into fuels for the end up in the gas stream [3,4]. Tar condenses at low temperatures
transportation sector and chemical production. Advanced ther- (<60  C) causing operational difficulties for the downstream pro-
mochemical technologies, i.e. pyrolysis and gasification, are cess (e.g. corrosion, clogging and fouling of installations) [4e6],
considered as potential approaches to reduce the heavy de- requiring tar treatment/removal steps. In addition, some carbon
pendency upon fossil fuels in the transportation sector (>90% from and hydrogen content of the feedstock is lost, thus reducing the
fossil based fuels) and their associated environmental impacts. energy efficiency of the process. Therefore, fully understanding the
Direct conversion of biomass into liquid fuel via pyrolysis is still effect of the pyrolysis on gasification stages is crucial to obtain high
encountering a number of issues due to thermally and chemically quality gas products in a gasification process.
unstable liquid products. However, biomass can be converted into Extensive research has been done on the pyrolysis of lignocel-
gas via gasification that can be either used for transportation fuels, lulosic materials, focusing on the effects of operating conditions on
i.e. H2 or converted further into hydrocarbons (via a Fischer- yields and properties of the process e.g. temperature [7e14],
Tropsch synthetic route) or methanol. The concept of gasification heating rate [15e19], vapour residence time [17,20e23] and parti-
of coal is well-known but there are issues when the same approach cle size of the feedstock [19,22,24e27]. However, these studies
is applied to biomass. This is because biomass consists of 75e85% investigated pyrolysis as a single process without providing any
volatile matter as compared to only 20e35% in coal. Large amounts links between the physicochemical properties of pyrolysis products
of volatiles in biomass are released rapidly in a very narrow and the operating parameters in the pyrolysis process that are
required for syngas/hydrogen production via a gasification route. In
this study, the effect of pyrolysis conditions on the gasification
process in terms of quality of synthetic gas (syngas) and tar for-
* Corresponding author. mation was investigated via two-stage gasification in which
E-mail address: anh.phan@ncl.ac.uk (A.N. Phan).

https://doi.org/10.1016/j.renene.2019.05.009
0960-1481/Crown Copyright © 2019 Published by Elsevier Ltd. All rights reserved.
P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120 113

feedstock was decomposed into intermediate products and sub- Table 1


sequently gasified, to provide an operating window for developing Physicochemical characteristics of waste wood.

a simple, highly efficient and robust gasification unit that can be Proximate analysis (wt%, dry basis) Values
used for green hydrogen production without major downstream Volatile matter 84.1 ± 0.4
processing to remove impurities. Fixed carbon 15.4 ± 0.1
Ash content 0.5 ± 0.1
Ultimate analysis (wt%, dry basis, ash free) Values
2. Materials and methods C 41.8 ± 0.3
H 6.4 ± 0.3
2.1. Feedstock Oa 51.5 ± 0.5
N 0.3 ± 0.3
HHV (MJ/kg) 17.7 ± 0.4
Waste wood obtained from Sustainable Campus, Newcastle Empirical formula of waste wood C6H11O5.5N0.04
University, was ground and sieved to 850 mm for characterisation a
By difference.
(Fig. 1a) and cut into cubes of 0.5 cm3, 1 cm3 and 2 cm3 using a
circular saw for experimental tests (Fig. 1b). The waste wood had a
very high volatile matter (84 wt%) and low ash content (<1 wt%) 20  C/min. As soon as the set temperature reached, i.e. 600, 700,
compared to coal (20e35 wt% volatile and 10e20 wt% ash) as 800 or 900  C, the system was held for a further 15 min before being
shown in Table 1. As a result, tighter controls are required in the switched off to ensure full decomposition of the volatiles released.
oxidation step in the gasification process to ensure that all volatiles The volatiles released from the pyrolysis were cooled down in two
have a good contact time with the oxidising agent in order to get condensers, kept in an ice bath (0  C), whereas the non-
high quality gas and a char product which is highly reactive for condensable gas was collected in a 500 ml tedlar sample bag for
solid-gas reactions occurring in a very short contact time (1e2 s) at gas analysis. The solid residues and condensable liquid were
high temperatures (>800  C), i.e. minimising the diffusion limita- collected and weighted for their yields and stored in glass bottles
tions to convert all the carbon content in the char into gas. Waste for further analysis when the reactor temperature was below 50  C.
wood had an empirical formula of C6H11O5.5N0.04 with 51 wt% ox- The experimental set-up to examine the effect of pyrolysis product
ygen content compared to around 7e10 wt% in coal. properties on the gasification stage is illustrated in Fig. 2b. The
volatiles released from the pyrolysis were passed through a char
2.2. Experimental procedure bed (which was produced by pyrolysis of waste wood at different
temperatures of 600, 700, 800 or 900  C as described previously)
Approximately 25.0 ± 0.1 g of the waste wood (size 0.5, 1 or heated at 1000  C. As soon as the volatiles from the pyrolysis stage
2 cm3) was placed in the centre of a 33 mm diameter and 830 mm were released (around 290e300  C), steam was injected at the T-
long Inconel 600 fixed bed (Fig. 2a). Prior to the experiments, the mixer before going into the gasification reactor at a steam to carbon
reactor was continuously purged with N2. As soon as the system in biomass (S/C) molar ratio of 5.7. The product collection and
was air free (confirmed by gas chromatography), the N2 was analysis procedure was the same as described previously in the
adjusted to a fixed flow rate of 30 or 120 ml/min to study the effect pyrolysis step.
of vapour residence time (1e3 s) on product yields and properties
in the pyrolysis step. The heating system was switched on at a fixed 2.3. Analysis
heating rate of 20  C/min; it was found in preliminary trials in this
study that there was no significant effect of the heating rate on Solid residues (char) were tested for their proximate analysis
product yields and characterisation of products between 5 and including moisture content, volatile matter and ash content based

(a) (b)
Fig. 1. Examples of waste wood samples used for (a) characterisation and (b) pyrolysis experiments (1 cm3 cube).
114 P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120

(a)

(b)
Fig. 2. Schematic of (a) pyrolysis experiment and (b) two-stage gasification experiment.

on a British standard (BS-1016-6). The physical morphology of the titration (915 KF Ti-Touch), according to the American Society for
waste wood sample before and after the pyrolysis was examined by Testing and Materials standard (ASTM E203-16). FT-IR analysis was
a Hitachi TM 3030 Scanning Electron Microscope (SEM) operated at carried out to characterize the organic functional groups in the
a 15 kV accelerating voltage. The specific surface area of char was liquid fraction using a ReactIR™ 4000 FT-IR spectrometer, with
determined from three random ground samples taken from a bulk scan range from 600 to 4000 cm-1. The chemical compositions of
sample by the Brunauer Emmett Teller (BET) nitrogen adsorption the liquid samples were identified and quantified using a 7200
technique using a Thermo Scientific Surfer Gas Adsorption Poros- Accurate-Mass Q-TOF GC/MS and gas chromatography flame ioni-
imeter. There was considerable point to point variability in the zation detector (GC-FID) equipped with a 60 m  250 mm x 0.25 mm
samples due to the expected variability of the natural feedstock. capillary column (14%-cyanopropyl-phenly-methylpolysiloxane,
The functional groups analysis of the char was carried out using an Restek Rtx-1707) with helium as carrier gas.
Agilent Cary 630 FT-IR Spectrometer with scan range from 700 to Non-condensable gas was analysed using a Varian 450-GC gas
4000 cm-1. The elemental analysis and the calorific value (HHV) of chromatograph (GC) equipped with 5 columns and 3 detectors (1
the char and condensed fraction (liquid) were carried out using an TCD and 2 FID) with argon as carrier gas. A Molecular sieve ultimate
Elementar Vario MICRO Cube CHN Elemental Analyser and a CAL2K 13 column (1.5 m  1/800  2.0 mm) was used to detect the per-
ECO Bomb calorimeter. manent gases (CO2, CO, H2, N2 and CH4). A Hayesep T and Q ulti-
A Metrohm 827 pH meter was used to measure the pH of the metal column (0.5 m  1/800  2.0 mm) was used to detect CO2, C2s
liquid (bio-oil) at atmospheric temperature (20  C). The water (acetylene, ethylene and ethane) and hydrocarbon gases. A CP-SIL
content in the liquid sample was measured using a Karl-Fischer 5CB capillary column (25 m  0.25 mm  0.4 mm) was used with a

Table 2
Pyrolysis product yields over various pyrolysis temperature and nitrogen flow rates at a fixed particle size of 1 cm3.

Temperature ( C)/N2 flow rate (ml/min) Char (wt%) Liquid (wt%) Gasa (wt%)

600/30 25.1 ± 0.2 50.5 ± 0.3 24.4 ± 0.2


600/120 25.3 ± 0.1 51.1 ± 0.1 23.6 ± 0.3
700/30 23.7 ± 0.1 50.5 ± 0.1 25.8 ± 0.2
700/120 23.7 ± 0.2 50.7 ± 0.3 25.6 ± 0.2
800/30 22.5 ± 0.3 47.7 ± 0.2 29.8 ± 0.1
800/120 22.5 ± 0.2 47.6 ± 0.2 29.9 ± 0.2
900/30 22.0 ± 0.2 46.9 ± 0.1 31.1 ± 0.1
900/120 22.2 ± 0.2 47.4 ± 0.2 30.4 ± 0.1
a
By difference.
P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120 115

Table 3
Properties of pyrolysis products from pyrolysis of waste wood at 600 700, 800 and 900  C at a fixed nitrogen flow rate of 120 ml/min and particle size of 1 cm3.

Temperature ( C) 600 700 800 900

Char properties (dry basis)


Volatile matter (wt%) 10.1 ± 2.2 7.2 ± 1.7 4.6 ± 1.4 3.6 ± 1.9
Fixed carbon (wt%) 87.5 ± 1.3 90.2 ± 2.1 92.7 ± 1.0 93.6 ± 1.0
Ash content (wt%) 2.4 ± 1.1 2.6 ± 0.7 2.7 ± 0.7 2.8 ± 0.3
C ± 0.3 (wt%) 85.8 87.0 87.3 87.6
H ± 0.3 (wt%) 2.4 1.9 1.7 1.5
O ± 0.5a (wt%) 11.4 10.6 10.5 10.3
N ± 0.3 (wt%) 0.4 0.5 0.5 0.6
The empirical formula C6H2O0.6N0.02 C6H1.6O0.5N0.03 C6H1.4O0.5N0.03 C6H1.2O0.5N0.04
HHV (MJ/kg) 32.5 ± 0.4 33.0 ± 0.1 33.5 ± 0.2 33.6 ± 0.4
BET surface area (m2/g) 38.6 ± 1.7 78.9 ± 3.9 82.0 ± 4.2 98.4 ± 4.6
Liquid properties (wet basis)
C ± 0.3 (wt%) 44.7 44.2 44.5 44.7
H ± 0.3 (wt%) 7.5 7.6 7.7 7.4
O ± 0.5a (wt%) 47.8 48.2 47.8 47.9
The empirical formula CH2O0.8 CH2.1O0.8 CH2.1O0.8 CH2O0.8
Water content in liquid fraction (wt%) 43.9 ± 0.4 44.1 ± 0.8 44.6 ± 0.5 43.7 ± 0.8
pH 2.3 ± 0.2 2.3 ± 0.1 2.4 ± 0.2 2.4 ± 0.3
HHV (MJ/kg) 17.3 ± 0.4 16.5 ± 0.3 17.5 ± 0.4 17.2 ± 0.2
Gas composition
H2 (mol%) 12.5 ± 0.1 16.3 ± 0.3 18.3 ± 0.2 22.3 ± 0.3
CO (mol%) 38.5 ± 0.2 42.7 ± 0.1 45.4 ± 0.1 47.6 ± 0.2
CO2 (mol%) 32.8 ± 0.1 27.5 ± 0.2 23.1 ± 0.1 15.1 ± 0.1
CH4 (mol%) 9.6 ± 0.3 11.3 ± 0.1 11.7 ± 0.1 13.4 ± 0.2
C2eC5 (mol%) 6.6 ± 0.4 2.2 ± 0.2 1.5 ± 0.2 1.6 ± 0.2
H2/CO 0.3 0.4 0.4 0.5
a
By difference.

FID to determine hydrocarbons. A CP-WAX 52CB capillary column decrease at pyrolysis temperatures above 550  C. Removal of
(25 m  0.32 mm  1.2 mm) equipped with another FID was used to functional groups at high temperatures provides char with high
detect light oxygenated compounds. stability and degree of condensation on the char [34e38].
The experimental results (Table 3) showed that the energy
content of the feedstock was mainly stored in the liquid and char.
3. Results and discussion

3.1. Effect of pyrolysis temperature on product yields and properties

Table 2 shows that operating temperatures of 600e700  C had


little effect on product yields. A further increase in temperature to
700e800  C resulted in a 23% increase in gas yield mainly at the
expense of the liquid. This is because the devolatilization of the
feedstock almost ceased at temperatures above 700  C and volatiles
were deeply cracked at high temperatures to form gases. This was
evidenced by the amount of fixed carbon and the char carbon
content remaining constant at 93e94 wt% and 87 wt% as shown in
Table 3. Although there is a small amount of volatiles (around
3e4 wt%) present in the char, they could be tightly bonded or
locked in stable structures and could not be released at the tested
operating conditions. The carrier gas flow rate in the range of
30e120 ml/min (volatiles residence time around 1e3 s) had little
effect on the product yields (Table 2). This could be because the
residence time was insufficient to promote significant cracking of
the volatiles. A small increase in the extent of cracking (around 4%)
has been reported by others at lower temperatures, i.e. 550  C
[13,17,19,23,28].
As observed in Fig. 3, hydroxyl groups (OH) at ~3300 cm-1 and
the aliphatic CeH stretching vibration at ~2900 cm-1 in the raw
material (Fig. 3a) disappeared at temperatures, i.e. 600  C (Fig. 3b),
whereas other functional groups, i.e. C]O (stretching vibration at
~1700 cm-1), aromatic CeC/C]C (vibration at ~1600 cm-1), CeO
(stretching (peak at ~1100 cm-1) and the aromatic CeH (stretch-
ing vibration between 700 cm-1 and 900 cm-1), decreased their
intensities with increasing pyrolysis temperatures and completely
disappeared at temperatures above 700  C (Fig. 3cee). The results Fig. 3. FT-IR spectra of (a) biomass feedstock (waste wood) and char obtained from
obtained from this study (Fig. 3) agree very well with other findings pyrolysis at a temperature of (b) 600  C; (c) 700  C; (d) 800  C and (e) 900  C at a fixed
[29e33], where functional groups on the char surface started to nitrogen flow rate of 120 ml/min and particle size of 1 cm3.
116 P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120

The energy yield (which is defined as calorific value of product*- temperature from 600 to 900  C. The average pore diameter of the
product yield/calorific value of raw material*mass of feedstock) was char surface was found to be in the range of 7e12 mm, suggesting a
around 39e46% in char and 42e47% in liquid over the tested range sufficient pore size to accommodate the volatiles from biomass, i.e.
of temperatures. oxygenated compounds, heterocyclic aromatic and polycyclic aro-
Fig. 4 shows the change in the char morphology with temper- matic hydrocarbons (PAHs) (<2 mm) [42e44]. High total pore and
ature. At a fixed particle size of 1 cm3, increasing temperature micropore volumes will promote the mass transfer between gas
accelerated the rate of decomposition [39e41], resulting in the and char particles, accelerating the heterogeneous solid-gas re-
formation of small pores on the surface of the char (where volatiles actions, thereby reducing unburnt carbon in the ash residues
are released) from the smooth solid cells in the raw material [45,46].
(Fig. 4a). The specific surface area of char increased from 38.6 m2/g Table 3 shows the CO2 concentration gradually decreased from
at 600  C to 98.4 m2/g at 900  C (Table 3). A significant increase in 32.8 to 15.1 mol%, whereas CO formation increased from 38.5 to
the total pore volume from 21.7  10-3 cm3/g to 52.1  10-3 cm3/g 47.6 mol% when increasing pyrolysis temperature from 600 to
and micropore (2e30 nm) volume from 13.1  10-3 cm3/g to 900  C. This is because when CO2 was released from inner layers of
34.4  10-3 cm3/g was observed when increasing the pyrolysis a particles it can react with a hot char layer via a Boudouard

500 μm

(a)

50 μm 50 μm

(b)

50 μm 50 μm

(c)
Fig. 4. SEM images of (a) waste wood raw material (at a magnification of 38) and char (at a magnification of 500) obtained from pyrolysis at a temperature of (b) 600  C and (c)
900  C at a fixed nitrogen flow rate of 120 ml/min and particle size of 1 cm3.
P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120 117

Table 4 almost zero oxygen) [52,53]. If all water is removed, the liquid
Chemical compositions of liquid derived from pyrolysis of waste wood over various product would have a H/C in a range of 1.3e1.4 and O/C of 0.4e0.5,
pyrolysis temperatures at a fixed nitrogen flow rate of 120 ml/min and particle size
of 1 cm3.
which is still higher than the required ratios for transport fuel ap-
plications. In addition, the liquid derived from pyrolysis is chemi-
Function groups wt% (dry basis) cally unstable due to more than 300 chemical compounds, oxygen-
600  C 700  C 800  C 900  C containing compounds in particular (i.e. hydroxyacetaldehyde, 5-
Acids 2.1 1.3 0.8 0.7 hydroxymethylfurfural, acetol, phenol and derivatives, levogluco-
Esters 1.4 1.6 0.5 0.6 san and anhydrosugars) [54e56] that are present in the liquid
Ketones 2.9 2.6 0.9 1.0 product. As shown in Table 4 full decomposition of unstable com-
Alcohols 1.1 1.1 0.8 0.7
pounds (i.e. sugars and their derivatives) was observed at pyrolysis
Aldehydes 1.1 1.1 0.6 0.6
Furan and its derivatives 4.9 4.4 4.9 3.7 temperatures above 700  C, while the concentration of other
Sugars 1.5 0.3 e e compounds derived from decomposition of cellulose and hemi-
Phenol and its derivatives 12.7 15.4 16.7 18.1 cellulose, i.e. acids, ketones, aldehydes and furans, gradually
Unknown 1.0 0.6 1.1 1.3 decreased with increasing temperature. However, phenol and its
derivatives increased with temperature due to Diels-Alder re-
actions which are favoured at high temperatures [57]. At all tested
reaction (C þ CO2 /2CO), which is dominant at temperatures temperatures the liquid consisted mainly of furan and its de-
above 700  C, leading to the formation of CO [47]. Moreover, CO2 rivatives and phenolic compounds (around 60% (Table 4)). This
could also react with other hydrocarbons (C1eC5) via a dry indicates that high temperature alone cannot decompose aromatics
   
reforming reaction Cx Hy þxCO2 42xCO þ 2y H2 to form H2 and into gaseous products.

CO at temperatures above 640oC [47,48]. Although, the H2/CO ratio


3.2. Effect of particle size on product yields and properties
increased when increasing temperature, i.e. from 0.3 at 600  C to
0.5 at 900  C due to an increase in H2 and CO (Table 3), it is still
Decreasing particle size from 2 cm3 to 0.5 cm3 had a significant
much lower than the required ratio for syngas applications, i.e.
effect on the char and gas yields, i.e. an approximately 31% decrease
chemical production and transportation fuel (H2/CO > 1) [49e51].
in the char yield and 29% increase in the gas yield (Table 5). This is
The liquid product was highly acidic with pH of 2.3e2.4 con-
due to heat transfer limitation in larger particles, causing higher
taining a large amount of water (around 44 wt% (based on liquid
temperature gradients inside the particles. It took around 46 min
fraction) or 22 wt% based on biomass feedstock (Table 3)). This
for the middle of 0.5 cm3 particle to reach the set temperature of
causes the corrosion of installations and difficulties in ignition due
900  C but 52 min for 1 cm3 and up to 62 min for 2 cm3. The outer
to low calorific value, thereby reducing the local combustion tem-
char layer could act as a hard shell hindering the release of volatiles
perature. The liquid product had much higher H/C and O/C ratios
[58,59], resulting in increasing char yield from 19.8 wt% for 0.5 cm3
(H/C~2.0 and O/C~0.8) than heavy fossil-based oil (H/C~1.5 and
to 28.7 wt% for 2 cm3) and the volatile content remaining in char

Table 5
Product yields and properties obtained from pyrolysis of waste wood at different particle sizes 0.5, 1 and 2 cm3 at pyrolysis temperature of 900  C and nitrogen flow rate of
120 ml/min.

Particle size (cm3) 0.5 1 2

Product yields
Char (wt%) 19.8 ± 0.2 22.2 ± 0.2 28.7 ± 0.2
Liquid (wt%) 47.6 ± 0.2 47.4 ± 0.2 46.0 ± 0.2
Gasa (wt%) 32.6 ± 0.2 30.4 ± 0.1 25.3 ± 0.1
Char properties (dry basis)
Volatile matter (wt%) 2.3 ± 1.0 3.6 ± 1.9 6.5 ± 1.4
Fixed carbon (wt%) 95.6 ± 2.2 93.6 ± 1.0 91.1 ± 1.9
Ash content (wt%) 2.1 ± 1.1 2.8 ± 0.3 2.4 ± 0.7
C ± 0.3 (wt%) 87.9 87.6 87.4
H ± 0.3 (wt%) 1.4 1.5 1.7
O ± 0.5a (wt%) 10.1 10.3 10.4
N ± 0.3 (wt%) 0.6 0.6 0.5
The empirical formula C6H1.1O0.5N0.04 C6H1.2O0.5N0.04 C6H1.4O0.5N0.03
HHV (MJ/kg) 34.0 ± 0.4 33.6 ± 0.4 33.3 ± 0.5
BET surface area (m2/g) 124.5 ± 5.3 98.4 ± 4.6 73.00 ± 3.2
Liquid properties (wet basis)
C ± 0.3 (wt%) 46.4 44.7 41.5
H ± 0.3 (wt%) 7.4 7.4 7.1
O ± 0.5a (wt%) 46.2 47.9 51.4
The empirical formula CH1.9O0.7 CH2O0.8 CH2O0.9
Water content in liquid fraction (wt%) 43.6 ± 0.4 43.7 ± 0.8 46.2 ± 0.4
pH 2.4 ± 0.2 2.4 ± 0.3 2.3 ± 0.3
HHV (MJ/kg) 18.0 ± 0.4 17.2 ± 0.2 14.9 ± 0.7
Gas composition
H2 (mol%) 24.4 ± 0.2 22.3 ± 0.3 18.2 ± 0.2
CO (mol%) 49.4 ± 0.3 47.6 ± 0.2 47.1 ± 0.3
CO2 (mol%) 11.0 ± 0.1 15.1 ± 0.1 20.5 ± 0.2
CH4 (mol%) 14.4 ± 0.3 13.4 ± 0.2 10.4 ± 0.2
C2eC5 (mol%) 0.8 ± 0.1 1.6 ± 0.2 3.8 ± 0.1
H2/CO 0.5 0.5 0.4
a
By difference.
118 P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120

50 μm 50 μm

(a) (b)

50 μm

(c)
Fig. 5. SEM images at a magnification of 500 of char obtained from (a) 0.5 cm ; (b) 1 cm3 and (c) 2 cm3 at pyrolysis temperature of 900  C and nitrogen flow rate of 120 ml/min.
3

(6.5 wt% for 2 cm3 but only 2.3 wt% for 0.5 cm3) as shown in Table 5. 0.5 cm3 to 14.9 MJ/kg at 2 cm3. Phenol and its derivatives decreased
Thus, understanding the temperature profile of single particles from 18.6 wt% at 0.5 cm3 to 13.7 wt% at 2 cm3 (Table 6). This could
could help in designing a gasifier that provides sufficient time for be due to an incomplete decomposition of lignin in the feedstock
decomposition of fuels. Although, there was little variation in C, H, [56,57,62]. However, the concentration of other compounds grad-
O content (Table 5), the specific surface area of the char was ually increased with increasing particle size as shown in Table 6.
significantly enhanced up to 124.5 m2/g at 0.5 cm3 compared to
73.0 m2/g at 2 cm3. The small variation in the repeated measure- 3.3. The effect of pyrolysis conditions on gasification process
ments (5%) was because different char samples derived from
different waste wood feedstock pieces were chosen when deter- Pyrolysis products obtained from three different pyrolysis
mining specific surface area. As shown in Fig. 5, the presence of temperatures (i.e. 600, 800 and 900  C) at a fixed particle size of
cracking and development of the pore structure (micropores and 1 cm3 and nitrogen flow rate of 120 ml/min were used to examine
macropores) was observed for the small particle size due to the the effect of pyrolysis on the syngas quality and tar formation in
rapid release of volatiles. Decreasing particle size from 2 cm3 to two-stage gasification. Gasification was fixed at a temperature of
0.5 cm3 led to increases in the total pore volume from 39.1  10- 1000  C and a steam to carbon in biomass (S/C) molar ratio of 5.7.
3
cm3/g to 66.3  10-3 cm3/g and the micropore volume from Table 7 shows that the tar concentration in the gas stream signifi-
23.1  10-3 cm3/g to 42.9  10-3 cm3/g. cantly decreased from 42.9 g/Nm3 (products derived at 600  C py-
Due to the significant difference in temperature observed for rolysis) to 21.2 g/Nm3 (products derived at 900  C pyrolysis),
large particles, an increase of CO2 from 11.0 to 20.5 mol% and light corresponding to approximately 50.6% tar removal. This could be
hydrocarbon (C2eC5) from 0.8 to 3.8 mol% was observed, whereas because the high specific surface area and pore volume of the char
other gas concentrations gradually decreased with increasing par- derived at high pyrolysis temperature (Table 3) promotes cracking/
ticle size from 0.5 cm3 to 2 cm3. As shown in Table 5, reducing reforming of tars [63e66]. Moreover, the high porosity of char also
particle size to 0.5 cm3 had little effect on the ratio of H2/CO enhances access of the gasifying agent (O2, CO2 or steam) to form
(around 0.5). H2 and CO via the heterogeneous solid-gas reactions occurs in the
As illustrated in Table 5, increasing particle size slightly gasification process, i.e. the Boudouard reaction (C þ CO2 /2CO)
decreased the carbon content in the liquid (from 46.4 wt% for and the water gas reaction (C þ H2 O/2CO þ H2 ) [67e69]. The H2
0.5 cm3 to 41.5 wt% for 2 cm3) but the water content in liquid and CO in the syngas increased from 48.8 to 67.2 mol% and
fraction increased slightly (around 6%). This could be due to the 4.5e8.8 mol% with the products derived at pyrolysis temperature of
rapid release of volatile matter, minimising secondary reactions (i.e. 600 and 900  C respectively (Table 7). Increasing the pyrolysis
dehydration) when small particles are used [13,14,60,61]. The temperature from 600 to 900  C alters the gas yield (from 77.8 to
calorific values (HHV) of the liquid decreased from 18.0 MJ/kg at 95.8 wt%) at the expense of solid residues (ash) (from 16.3 to 0.4 wt
P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120 119

Table 6 0.5 cm3. There were strong interactions between pyrolysis oper-
Compositions of liquid (condensable volatiles) derived from pyrolysis of waste wood ating conditions and the gasification step. Temperature in the py-
over a range of particle sizes (0.5, 1 and 2 cm3) at pyrolysis temperature of 900  C
and nitrogen flow rate of 120 ml/min.
rolysis process strongly altered the char morphology (around 2e3
time increase in surface areas, total pore and micropore volumes)
Functional groups wt% (dry basis) and composition of the volatile products, which in turn had a
0.5 cm3 1 cm3 2 cm3 strong influence on the syngas properties, i.e. tar, carbon in ash
Acids 0.7 0.7 0.9 (residues) and hydrogen content in gasification. Increasing pyrol-
Esters 1.6 0.6 1.1 ysis temperature increased hydrogen content whilst decreasing
Ketones 0.8 1.0 2.2 hydrocarbons, CO2 and tar content in the gas stream. Pyrolysis
Alcohols 1.0 0.7 1.3
temperatures above 700  C promoted the formation of phenolic
Aldehydes 0.6 0.6 1.0
Furan and its derivative 2.8 3.7 4.3 compounds that would be precursors of multiple aromatic ring
Phenol and its derivatives 18.6 18.1 13.7 species in the gas products in gasification. Nonetheless, only
Unknown 0.8 1.3 0.3 decreasing size to 0.5 cm3 and applying high temperature in py-
rolysis would not be sufficient to produce high H2/CO ratio (only
0.4e0.5). Our findings show that at a pyrolysis temperature of
Table 7 900  C, a particle size of 0.5e1 cm3 and at a controlled steam to
The effect of pyrolysis temperature on two-stage gasification at a fixed gasification carbon in biomass (S/C) molar ratio, i.e. 5.7, the H2 content in
temperature of 1000  C and a steam to carbon in biomass (S/C) molar ratio of 5.7.
gasification products increased up to 67 mol% from around 49%
Pyrolysis temperature ( C) 600 800 900 whereas the aromatic compounds decreased by up to 50.6%. The
Gas yield (wt%) 77.8 ± 3.4 88.0 ± 2.1 95.8 ± 3.1 carbon content in ash (residues) from gasification was less than
CO2 (mol%) 39.0 ± 2.3 31.9 ± 1.2 21.8 ± 3.3
0.01 wt%.
H2 (mol%) 48.8 ± 1.3 56.8 ± 1.2 67.2 ± 1.4
CH4 (mol%) 7.7 ± 0.7 4.7 ± 1.2 2.2 ± 1.2
CO (mol%) 4.5 ± 0.8 6.6 ± 1.1 8.8 ± 0.3
Tar (g/Nm3) 42.9 34.7 21.2 References
Solid residues (wt%) 16.3 ± 1.4 7.3 ± 0.7 0.4 ± 1.1
[1] D.K. Seo, S.S. Park, J. Hwang, T.-U. Yu, Study of the pyrolysis of biomass using
thermo-gravimetric analysis (TGA) and concentration measurements of the
evolved species, J. Anal. Appl. Pyrolysis 89 (1) (2010) 66e73.
%) and unburnt carbon in the ash residues (<0.01 wt%) at a fixed [2] A. Soria-Verdugo, E. Goos, N. García-Hernando, Effect of the number of TGA
curves employed on the biomass pyrolysis kinetics results obtained using the
gasification temperature of 1000  C and a steam to carbon in Distributed Activation Energy Model, Fuel Process. Technol. 134 (2015)
biomass (S/C) molar ratio of 5.7 (Table 7). Moreover, an enhance- 360e371.
ment in the gas properties (higher CO concentration in the gas [3] U. Arena, Process and technological aspects of municipal solid waste gasifi-
cation. A review, Waste Manag. 32 (4) (2012) 625e639.
stream), which results in an increased pyrolysis temperature [4] C. Li, K. Suzuki, Tar property, analysis, reforming mechanism and model for
(Table 3), promotes a significant reaction, i.e. water gas shift reac- biomass gasificationdan overview, Renew. Sustain. Energy Rev. 13 (3) (2009)
tion ðCO þ H2 O 4CO2 þ H2 Þ, which occurs in the gasification step 594e604.
[5] T. Milne, E.R.a.A., Biomass Gasifier “Tars”: Their Nature, Formation and Con-
leading to the formation of H2 in the gas stream.
version, National Renewable Energy Laboratory, 1998.
Parameters affecting the pyrolysis process and the gas proper- [6] L. Zhang, C. Xu, P. Champagne, Overview of recent advances in thermo-
ties derived from gasification in this study can be applied for other chemical conversion of biomass, Energy Convers. Manag. 51 (5) (2010)
high-volatile containing feedstock (75e85%). The initial waste 969e982.
[7] A.P.a.K. Yoshikawa, Influence of pyrolysis temperature on rice husk char
wood composition (C, H, N, O) was found to have little influence on characteristics and its tar adsorption capability, Energies 5 (2012) 4941e4951.
products derived from pyrolysis process [70e72]. However, the [8] M.A.F. Mazlan, Y. Uemura, N.B. Osman, S. Yusup, Fast pyrolysis of hardwood
amount of steam required for the gasification process should be residues using a fixed bed drop-type pyrolyzer, Energy Convers. Manag. 98
(2015) 208e214.
adjusted according to the initial carbon content in the feedstock to [9] Z. Yang, A. Kumar, R.L. Huhnke, M. Buser, S. Capareda, Pyrolysis of eastern
maintain the optimum ratio. Ash in lignocellulosic materials con- redcedar: distribution and characteristics of fast and slow pyrolysis products,
tains mainly alkaline and other metals (Na, K, Si, Ca, Mg etc.) and Fuel 166 (2016) 157e165.
[10] N. Bordoloi, R. Narzari, D. Sut, R. Saikia, R.S. Chutia, R. Kataki, Characterization
remains as a solid residue. These metals could act as catalysts for of bio-oil and its sub-fractions from pyrolysis of Scenedesmus dimorphus,
reforming/cracking of high molecular weight compounds in both Renew. Energy 98 (2016) 245e253.
pyrolysis and gasification processes [64,65], promoting less tar [11] A. Shaaban, S.-M. Se, N.M.M. Mitan, M.F. Dimin, Characterization of biochar
derived from rubber wood sawdust through slow pyrolysis on surface po-
formation. However, these inorganic species could cause a number rosities and functional groups, Procedia Eng. 68 (2013) 365e371.
of challenges such as fouling, erosion and corrosion of installations [12] J. Akhtar, N.A.S. Amin, A review on process conditions for optimum bio-oil
as well as slagging and agglomeration when the gasification tem- yield in hydrothermal liquefaction of biomass, Renew. Sustain. Energy Rev.
15 (3) (2011) 1615e1624.
perature is above 1200  C (melting point). The process developed in
[13] T. Aysu, M.M. Küçük, Biomass pyrolysis in a fixed-bed reactor: effects of py-
this study is carried out at a temperature range of 1000e1100  C, rolysis parameters on product yields and characterization of products, Energy
which only just enough to transfer these metals into stable forms, 64 (2014) 1002e1025.
i.e. NaO, Al2O3, P2O5, K2O, K2CO3, KCl, SiO2, Fe2O3, CaCO3, CaO and [14] Y.L. Tan, A.Z. Abdullah, B.H. Hameed, Fast pyrolysis of durian (Durio zibethinus
L) shell in a drop-type fixed bed reactor: pyrolysis behavior and product
MgO [73,74]. analyses, Bioresour. Technol. 243 (2017) 85e92.
[15] D. Chen, J. Zhou, Q. Zhang, Effects of heating rate on slow pyrolysis behavior,
kinetic parameters and products properties of moso bamboo, Bioresour.
4. Conclusions Technol. 169 (2014) 313e319.
[16] D. Chen, Y. Li, K. Cen, M. Luo, H. Li, B. Lu, Pyrolysis polygeneration of poplar
The effects of pyrolysis temperature (600e900  C), nitrogen wood: effect of heating rate and pyrolysis temperature, Bioresour. Technol.
218 (2016) 780e788.
flow rate (30 and 120 ml/min corresponding to volatile residence €z, Pyrolysis of hornbeam (Carpinus betulus L.)
[17] U. Moralı, N. Yavuzel, S. Şenso
time (1e3 s) and particle size (0.5, 1 and 2 cm3) on the product sawdust: characterization of bio-oil and bio-char, Bioresour. Technol. 221
yields and properties of pyrolysis of waste wood were investigated. (2016) 682e685.
Decreasing the particle size of the feedstock increased the specific [18] S. Yorgun, D. Yıldız, Slow pyrolysis of paulownia wood: effects of pyrolysis
parameters on product yields and bio-oil characterization, J. Anal. Appl. Py-
surface area and total pore volume of the char from 73.0 m2/g and rolysis 114 (2015) 68e78.
39.1  10-3 cm3/g at 2 cm3 to 124.5 m2/g and 66.3  10-3 cm3/g at [19] A.K. Varma, P. Mondal, Pyrolysis of sugarcane bagasse in semi batch reactor:
120 P. Prasertcharoensuk et al. / Renewable Energy 143 (2019) 112e120

effects of process parameters on product yields and characterization of J. Anal. Appl. Pyrolysis 118 (2016) 42e53.
products, Ind. Crops Prod. 95 (2017) 704e717. [46] E. Arenas, F. Chejne, The effect of the activating agent and temperature on the
[20] S.A. Jourabchi, S. Gan, H.K. Ng, Pyrolysis of Jatropha curcas pressed cake for porosity development of physically activated coal chars, Carbon 42 (12)
bio-oil production in a fixed-bed system, Energy Convers. Manag. 78 (2014) (2004) 2451e2455.
518e526. [47] E.E. Kwon, Y.J. Jeon, H. Yi, New candidate for biofuel feedstock beyond
[21] H.V. Ly, S.-S. Kim, J.H. Choi, H.C. Woo, J. Kim, Fast pyrolysis of Saccharina terrestrial biomass for thermo-chemical process (pyrolysis/gasification)
japonica alga in a fixed-bed reactor for bio-oil production, Energy Convers. enhanced by carbon dioxide (CO2), Bioresour. Technol. 123 (2012) 673e677.
Manag. 122 (2016) 526e534. [48] A.M.M. Leal, D.A. Kulik, G. Kosakowski, Computational methods for reactive
[22] J.M. Encinar, J.F. Gonza lez, J. Gonz
alez, Fixed-bed pyrolysis of Cynara car- transport modeling: a Gibbs energy minimization approach for multiphase
dunculus L. Product yields and compositions, Fuel Process. Technol. 68 (3) equilibrium calculations, Adv. Water Resour. 88 (2016) 231e240.
(2000) 209e222. [49] S.T. Chaudhari, S.K. Bej, N.N. Bakhshi, A.K. Dalai, Steam gasification of biomass-
[23] U. Moralı, S. Şenso € z, Pyrolysis of hornbeam shell (Carpinus betulus L.) in a derived char for the production of carbon monoxide-rich synthesis gas, En-
fixed bed reactor: characterization of bio-oil and bio-char, Fuel 150 (2015) ergy Fuels 15 (3) (2001) 736e742.
672e678. [50] W. Torres, S.S. Pansare, J.G. Goodwin, Hot gas removal of tars, ammonia, and
[24] W.K. Buah, A.M. Cunliffe, P.T. Williams, Characterization of products from the hydrogen sulfide from biomass gasification gas, Catal. Rev. 49 (4) (2007)
pyrolysis of municipal solid waste, Process Saf. Environ. Protect. 85 (5) (2007) 407e456.
450e457. [51] H. Yang, R. Yan, H. Chen, D.H. Lee, C. Zheng, Characteristics of hemicellulose,
[25] J. Shirley, J.L. Duarte, Dario Alviso, Juan C. Rolon, Effect of temperature and cellulose and lignin pyrolysis, Fuel 86 (12) (2007) 1781e1788.
particle size on the yield of bio-oil, produced from conventional coconut core [52] L. Zhang, R. Liu, R. Yin, Y. Mei, Upgrading of bio-oil from biomass fast pyrolysis
pyrolysis, Int. J. Chem. Eng. Appl. 7 (2) (2016) 102e108. in China: a review, Renew. Sustain. Energy Rev. 24 (2013) 66e72.
[26] S. Şenso _ Demiral, H. Ferdi Gerçel, Olive bagasse (Olea europea L.) pyrolysis,
€ z, I. [53] A. Oasmaa, S. Czernik, Fuel oil quality of biomass pyrolysis oils state of the art
Bioresour. Technol. 97 (3) (2006) 429e436. for the end users, Energy Fuels 13 (4) (1999) 914e921.
[27] C. Di Blasi, C. Branca, V. Lombardi, P. Ciappa, C. Di Giacomo, Effects of particle [54] A. Oasmaa, E. Leppa €ma €ki, P. Koponen, J. Levander, E. Tapola, Physical Char-
size and density on the packed-bed pyrolysis of wood, Energy Fuels 27 (11) acterisation of Biomass-Based Pyrolysis Liquids Application of Standard Fuel
(2013) 6781e6791. Oil Analyses, 1997.
[28] H.F. Gerçel, Production and characterization of pyrolysis liquids from [55] A.O. Jani Lehto, Yrjo € Solantausta, Matti Kyto € , David Chiaramonti, Fuel Oil
sunflower-pressed bagasse, Bioresour. Technol. 85 (2) (2002) 113e117. Quality and Combustion of Fast Pyrolysis Bio-Oils, 2013.
[29] W.-K. Kim, T. Shim, Y.-S. Kim, S. Hyun, C. Ryu, Y.-K. Park, J. Jung, Character- [56] J. Alvarez, G. Lopez, M. Amutio, J. Bilbao, M. Olazar, Bio-oil production from
ization of cadmium removal from aqueous solution by biochar produced from rice husk fast pyrolysis in a conical spouted bed reactor, Fuel 128 (2014)
a giant Miscanthus at different pyrolytic temperatures, Bioresour. Technol. 162e169.
138 (2013) 266e270. [57] R. Cypres, Aromatic hydrocarbons formation during coal pyrolysis, Fuel Pro-
[30] W. Chen, S. Shi, T. Nguyen, M. Chen, X. Zhou, Effect of Temperature on the cess. Technol. 15 (1987) 1e15.
Evolution of Physical Structure and Chemical Properties of Bio-Char Derived [58] H. Bennadji, K. Smith, M.J. Serapiglia, E.M. Fisher, Effect of particle size on low-
from Co-pyrolysis of Lignin with High-Density Polyethylene, vol. 11, 2016, temperature pyrolysis of woody biomass, Energy Fuels 28 (12) (2014)
pp. 3923e3936. 7527e7537.
[31] Y. Elmay, Y. Le Brech, L. Delmotte, A. Dufour, N. Brosse, R. Gadiou, Charac- [59] S. Hanson, J.W. Patrick, A. Walker, The effect of coal particle size on pyrolysis
terisation of miscanthus pyrolysis by DRIFTs, UV Raman Spectrosc. Mass and steam gasification, Fuel 81 (5) (2002) 531e537.
Spectrom. 113 (2015). [60] X. Bai, P. Johnston, S. Sadula, R.C. Brown, Role of levoglucosan physi-
[32] W.-J. Liu, H. Jiang, H.-Q. Yu, Development of biochar-based functional mate- ochemistry in cellulose pyrolysis, J. Anal. Appl. Pyrolysis 99 (2013) 58e65.
rials: toward a sustainable platform carbon, Mater. Chem. Rev. 115 (22) [61] F. Ronsse, S. van Hecke, D. Dickinson, W. Prins, Production and characteriza-
(2015) 12251e12285. tion of slow pyrolysis biochar: influence of feedstock type and pyrolysis
[33] X. Gai, H. Wang, J. Liu, L. Zhai, S. Liu, T. Ren, H. Liu, Effects of feedstock and conditions, GCB Bioenergy 5 (2) (2013) 104e115.
pyrolysis temperature on biochar adsorption of ammonium and nitrate, PLoS [62] C. Branca, P. Giudicianni, C. Di Blasi, GC/MS characterization of liquids
One 9 (12) (2014), e113888. generated from low-temperature pyrolysis of wood, Ind. Eng. Chem. Res. 42
[34] M.K. Rafiq, R.T. Bachmann, M.T. Rafiq, Z. Shang, S. Joseph, R. Long, Influence of (14) (2003) 3190e3202.
pyrolysis temperature on physico-chemical properties of corn stover (Zea [63] A. Paethanom, K. Yoshikawa, Influence of pyrolysis temperature on rice husk
mays L.) biochar and feasibility for carbon capture and energy balance, PLoS char characteristics and its tar adsorption capability, Energies 5 (12) (2012)
One 11 (6) (2016), e0156894. 4941.
[35] A. Enders, K. Hanley, T. Whitman, S. Joseph, J. Lehmann, Characterization of [64] Y. Shen, Y. Fu, Advances in in situ and ex situ tar reforming with biochar
biochars to evaluate recalcitrance and agronomic performance, Bioresour. catalysts for clean energy production, Sustain. Energy Fuels 2 (2) (2018)
Technol. 114 (2012) 644e653. 326e344.
[36] N.A.d. Figueredo, L.M.d. Costa, L.C.A. Melo, E.A. Siebeneichlerd, J. Tronto, [65] J. Liu, Y. He, X. Ma, G. Liu, Y. Yao, H. Liu, H. Chen, Y. Huang, C. Chen, W. Wang,
Characterization of biochars from different sources and evaluation of release Catalytic Pyrolysis of Tar Model Compound with Various Bio-Char Catalysts to
of nutrients and contaminants, Rev. Cienc. Agron. 48 (2017) 3e403. Recycle Char from Biomass Pyrolysis, vol. 11, 2016, pp. 3752e3768.
[37] W. Song, M. Guo, Quality variations of poultry litter biochar generated at [66] S. Anis, Z.A. Zainal, Tar reduction in biomass producer gas via mechanical,
different pyrolysis temperatures, J. Anal. Appl. Pyrolysis 94 (2012) 138e145. catalytic and thermal methods: a review, Renew. Sustain. Energy Rev. 15 (5)
[38] K. Jindo, H. Mizumoto, Y. Sawada, M. Sa nchez-Monedero, T. Sonoki, Physical (2011) 2355e2377.
and Chemical Characterizations of Biochars Derived from Different Agricul- [67] S. Ren, H. Lei, L. Wang, Q. Bu, S. Chen, J. Wu, Hydrocarbon and hydrogen-rich
tural Residues, vol. 11, 2014. syngas production by biomass catalytic pyrolysis and bio-oil upgrading over
[39] D. Pradhan, R.K. Singh, H. Bendu, R. Mund, Pyrolysis of Mahua seed (Madhuca biochar catalysts, RSC Adv. 4 (21) (2014) 10731e10737.
indica) e production of biofuel and its characterization, Energy Convers. [68] J.A. Mene ndez, A. Domínguez, Y. Ferna ndez, J.J. Pis, Evidence of self-
Manag. 108 (2016) 529e538. gasification during the microwave-induced pyrolysis of coffee hulls, Energy
[40] M. Ertaş, M. Hakkı Alma, Pyrolysis of laurel (Laurus nobilis L.) extraction Fuels 21 (1) (2007) 373e378.
residues in a fixed-bed reactor: characterization of bio-oil and bio-char, [69] Y. Shen, Chars as carbonaceous adsorbents/catalysts for tar elimination during
J. Anal. Appl. Pyrolysis 88 (1) (2010) 22e29. biomass pyrolysis or gasification, Renew. Sustain. Energy Rev. 43 (2015)
[41] T. Abdel-Fattah, M.E. Mahmoud, S.B. Ahmed, M. Huff, J. Lee, S. Kumar, Biochar 281e295.
from Woody Biomass for Removing Metal Contaminants and Carbon [70] P. Fu, W. Yi, X. Bai, Z. Li, S. Hu, J. Xiang, Effect of temperature on gas
Sequestration, vol. 22, 2014. composition and char structural features of pyrolyzed agricultural residues,
[42] H.-H. Yang, S.-M. Chien, M.-R. Chao, C.-C. Lin, Particle size distribution of Bioresour. Technol. 102 (17) (2011) 8211e8219.
polycyclic aromatic hydrocarbons in motorcycle exhaust emissions, J. Hazard. [71] Z. Luo, S. Wang, Y. Liao, J. Zhou, Y. Gu, K. Cen, Research on biomass fast py-
Mater. 125 (1) (2005) 154e159. rolysis for liquid fuel, Biomass Bioenergy 26 (5) (2004) 455e462.
[43] S.P. Wu, S. Tao, W.X. Liu, Particle size distributions of polycyclic aromatic [72] R. Xu, L. Ferrante, C. Briens, F. Berruti, Flash pyrolysis of grape residues into
hydrocarbons in rural and urban atmosphere of Tianjin, China, Chemosphere biofuel in a bubbling fluid bed, J. Anal. Appl. Pyrolysis 86 (1) (2009) 58e65.
62 (3) (2006) 357e367. [73] V. Benedetti, F. Patuzzi, M. Baratieri, Characterization of char from biomass
[44] R. Ladji, N. Yassaa, C. Balducci, A. Cecinato, Particle size distribution of n-al- gasification and its similarities with activated carbon in adsorption applica-
kanes and polycyclic aromatic hydrocarbons (PAHS) in urban and industrial tions, Appl. Energy 227 (2018) 92e99.
aerosol of Algiers, Algeria, Environ. Sci. Pollut. Control Ser. 21 (3) (2014) [74] S.V. Vassilev, D. Baxter, L.K. Andersen, C.G. Vassileva, An overview of the
1819e1832. composition and application of biomass ash. Part 1. Phaseemineral and
[45] D. Rehrah, R.R. Bansode, O. Hassan, M. Ahmedna, Physico-chemical charac- chemical composition and classification, Fuel 105 (2013) 40e76.
terization of biochars from solid municipal waste for use in soil amendment,

You might also like