Sei sulla pagina 1di 11

Biomass and Bioenergy 119 (2018) 335–345

Contents lists available at ScienceDirect

Biomass and Bioenergy


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

Research paper

Gasification of lignocellulosic biomass to produce syngas in a 50 kW T


downdraft reactor
Minhaj Uddin Monira,b, Azrina Abd Aziza,∗, Risky Ayu Kristantia, Abu Yousufc
a
Faculty of Engineering Technology, Universiti Malaysia Pahang, 26300, Gambang, Malaysia
b
Department of Petroleum and Mining Engineering, Jessore University of Science and Technology, Jessore, 7408, Bangladesh
c
Department of Chemical Engineering and Polymer Science, Shahjalal University of Science and Technology, Sylhet, 3114, Bangladesh

ARTICLE INFO ABSTRACT

Keywords: Lignocellulosic biomass gasification shows a pronounced prospective to replace fossil fuels. In this study, the
Co-gasification gasification of coconut shell with charcoal using a 50 kW downdraft reactor was investigated. The controlling
Coconut shell parameter of temperature and pressure were used to verify the production of gas during the gasification process
Charcoal with air. The higher contents of cellulose and hemicellulose than lignin in the sample were found to gasify better,
Downdraft reactor
as evident from structural analysis. The gasifier produces a combustible gas with a H2, CO, CO2 and CH4 con-
Syngas
centrations of 8.44, 15.38, 5.38 and 1.62 mol.% respectively, at a total flow of air of 30 m3 h−1. The results
revealed that 30 wt% charcoal in the feedstock was effectively gasified to generate syngas comprising over
30 mol.% of syngas with a lower heating value of 3.27 MJ/Nm3. Thus, the co-gasification of lignocellulosic
biomass with charcoal may contribute to affordable and environmentally friendly syngas energy.

1. Introduction four separate thermochemical conversion zones. These zones are


drying, pyrolysis, oxidation and reduction, included from top to
Worldwide energy demand is increasing exponentially due to the bottom, respectively, while reduction zone is responsible for syngas
rising trend of global population, industrial expansion, rapid urbani- production. This type of reactor produces less amount of tar, and re-
zation, economic growth which results the gradual depletion of fossil- duction zone plays a vital role for the maximum conversion of tar and
based fuels like oil, gas and coal [1–3]. Lignocellulosic biomass is an high quality syngas [11]. Syngas consists of H2, CO, CO2, CH4, N2 and
alternative, promising and fastest-growing renewable energy sources to some other compounds like C2H4, C2H6, NH3, H2S, and tar [9,12]. The
produce bioenergy [4,5]. This type of biomass absorbs and stored en- production of syngas is affected by some important factors such as re-
ergy from the sunlight in the form of chemical energy through photo- actor type, fuel type, gasification medium (steam or air) and opera-
synthesis process. These biomasses contain 40 to 50 wt% of cellulose, tional conditions (temperature, pressure) whereas fuel type has the
25 to 35 wt% of hemicellulose and 16 to 33 wt% of lignin [6]. Gen- most significant impact. The previous work stated that the co-gasifi-
erally, biomass consists of chemical bonds of CeC, CeO and some other cation rate in a downdraft reactor was increased gradually due to the
elements of nitrogen (N) and sulfur (S) [7]. Bioenergy containing bio- addition of char (charcoal) with coconut shell [13].
mass is converted into energy through thermochemical conversion of The co-gasification process was performed for the reduction of tar
combustion, pyrolysis and gasification [4]. Gasification is the thermo- content in the produced syngas by blending of biomass with coal [14].
chemical partial oxidation process that converted biomass into syn- Some researchers reported the co-gasification process using fossil-based
thetic gas (syngas) using a reactor and further processing it is used for fuel (coal) and lignocellulosic biomass considering the environmental
electricity or transporting fuel purposes [8]. Generally, there are four and gasification enhancement point of view [15–19]. Furthermore,
types of reactors (downdraft, updraft, fluidized bed and entrained bed) coal-based energy is treated as a dirty energy due to the presence of
are used for syngas production through this conversion process. sulfur, contrariwise lignocellulosic biomass and biomass-based charcoal
Biomass gasification through downdraft reactor is most widely used contain least amount of sulfur. In addition, exploitation, processing and
due to its high conversion efficiency of the producer syngas [9]. The transportation cost of coal is relatively high [20,21]. Collard and Blin
choice of reactor depends on feedstock type, size, moisture content, and [22] reported that conversion of biomass-based feedstock (cellulose,
gasification agents (steam/oxygen/air) [10]. Downdraft reactor has hemicelluloses and lignin) through thermochemical process followed


Corresponding author.
E-mail address: azrinaaziz@ump.edu.my (A. Abd Aziz).

https://doi.org/10.1016/j.biombioe.2018.10.006
Received 20 March 2018; Received in revised form 28 September 2018; Accepted 3 October 2018
0961-9534/ © 2018 Elsevier Ltd. All rights reserved.
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

three main pathways (char formation, depolymerization and fragmen- samples of approximately, 2 mg were placed in a tin capsule and
tation) and some reactions. The lignin contains various aromatic com- crimped and were placed in the auto sampler. The oxidation tempera-
pounds that convert into highest char products, and through depoly- ture was set as 1000 °C. The carrier of helium (He) gas was used for this
merization, it is further converted into some high molecular weight analysis. The analytical elemental composition of all samples was given
compounds. The fragmentation of various compounds can be changed in percentage (%).
into CO, CO2 and some small organic compounds. Due to high tem- The fourier transform infrared spectroscopy (FTIR) analysis was
perature (> 500 °C), pyrolyzed residues (charcoal) can be produced as achieved by detecting the functional group of coconut shell, charcoal
condensed polyaromatic compounds by releasing gaseous products of and co-product tar using a spectrometer (Perkin–Elmer, 670 FTIR). The
CO, H2 and CH4 [22]. Therefore, byproduct charcoal blending with ratio between sample and potassium bromide (KBr) was about 1:10
coconut shell could provide a promising opportunity by performing a with the rate of 0.5 cm/s. The FTIR spectra were investigated in a
co-gasification process using a pilot scale downdraft reactor. transmission mode and scanned over the wavenumbers range from
In the previous study, temperature and pressure of the reactor were 400 cm−1 to 4000 cm−1 with a resolution of 4 cm−1. All spectral sets of
optimized by using an aspen plus (V.8.6) simulator and enhanced the data were studied by using Principle Component Analysis (PCA) to
gasification process by adding charcoal to the coconut shell in a identify the chemical variations [26].
downdraft reactor. The optimized conditions reduce the overall ex- The X-ray diffraction (XRD) analysis was done by X-ray dif-
perimental cost. Very limited works have been done on the co-gasifi- fractometer (XRD-6000, Shimadzu) and Cu-Kαradiation (wave length,
cation of coconut shell and charcoal. Syngas production from biomass λ = 0.15418 nm) was used as the source. In this analysis, the coconut
have been studied by some researchers [15,23], and synergy, char- shell and charcoal were characterized to understand the crystalline
acterization and yield gas composition were focused mainly. In addi- characteristics of the material. The X-ray diffraction patterns were set
tion, the effect of carbon monoxide (CO) on the co-gasification process for the 2θ range of 10–80° with 2θ of 5° per minute. The crystal di-
using coal and biomass has been outlined recently [16]. In literature, to mension was calculated by using Scherrer equation [27].
the best knowledge of authors, characterization of lignocellulosic bio- The transmission electron microscopy (TEM) analysis was per-
mass of coconut shell, and experimental work through the co-gasifica- formed by examining the morphological structure (size, shape, reg-
tion process using coconut shell with various ratios of charcoal in a ularity and agglomeration of nanoparticles and cellulose) of coconut
downdraft reactor has not been studied sufficiently yet. The objective of shell and charcoal using a transmission electron microscope (FEI,
this study has two parts, first to characterize the feedstocks of coconut Tecnai-G2-20-Twin), operated at an acceleration voltage of 200 K. The
shell and charcoal for their bioenergy potentiality while calculating the co-product tar was characterized by scanning electron microscopy with
lower heating value (LHVbiomass ), higher heating value (HHVbiomass ), energy dispersive X-ray (SEM with EDX) spectroscopy (FEI, Quanta
exergy for feedstock (Exfeedstocks ). In the second part, the exergy value of 450) to identify the elemental and morphological features. Tar com-
produced syngas is compared with the characterization and experi- pounds were observed by gas chromatography-mass spectrometry (GC-
mental results obtained from the co-gasification of lignocellulosic bio- MS) analyzer (Agilent, 7890A) and nuclear magnetic resonance (1H
mass of coconut shell and charcoal using the downdraft reactor. NMR and 13C NMR) analyzer (Bruker, Bruker Ultra Shield Plus
500MHz/400 MHz). These analyses were carried out to identify the
2. Materials and methods nature and type of organic compounds in the co-product tar. During GC-
MS analysis, the initial and final temperatures were set as 70 °C and
2.1. Feedstock preparation and compositional analysis 325 °C, respectively. The heater temperature was set as 300 °C, pressure
was 8.8085 psi, and flow rate was programmed as 1 mL/min for 0 min.
In this co-gasification study, coconut shell and byproduct charcoal The MS was operated in the scan mode mass ranges from 40 amu to 600
were selected as the feedstocks. The coconut fruit was collected from amu. The X-ray photoelectron spectroscope (XPS) analysis was used to
the local fruit market near Gambang, Pahang, Malaysia. It was scrapped observe the chemical composition and its binding energies for coconut
by separating fibres (husk) and removed inner fruit from the shell. shell and charcoal using a PHI 5000 VersaProbeII Scanning X-ray
Then, coconut shell was crushed by a hand hammer. Subsequently, it photoelectron spectroscope (XPS) Microprobe.
was kept under the sun (two weeks) to maintain its dryness and im- The produced syngas was examined by an online gas chromato-
mediately placed in the sample bag to avoid any contamination. graphy-thermal conductivity detector (GC-TCD) (GC-2014, Shimadzu).
Charcoal was also collected from previous biomass gasification. The analyzed gases were mainly nitrogen (N2), hydrogen (H2), carbon
Subsequently, these feedstocks were cut, chipped and crushed as per dioxide (CO2), carbon monoxide (CO), methane (CH4), etc. The carrier
required size (10 mm–20 mm) for downdraft reactor [24,25]. More- gas of helium (He) was used for this analysis.
over, the desired particle size for coconut shell and charcoal were There are a number of equations (Eqs. (1)–(4)) used for the calcu-
prepared as less than 300 μm for characterization. Then, samples were lation of heating value (for biomass and produced gas) studied in the
also dried at 55 °C in a furnace for 72 h. Finally, samples were reserved literature based on the proximate and ultimate analytical results
in a tight screw cap plastic bottles for characterization. [28–31]:
The thermal decomposition of both feedstock was investigated by HHVbiomass = 0.3491MC + 1.1783MH + 0.1005MS 0.1034MO 0.0151MN 0.0211MAC
proximate analysis (Mettler Toledo, TGA/DSC1) and calculated the
(1)
relative percentage of moisture content (MC), ash content (AC) volatile
matter (VM) and fixed carbon (FC) reported in the literature [24].
Prepared powder sample (approx. 10 mg) of each coconut shell and
Where, HHVbiomass = The higher heating value of biomass ( );
MJ
kg

MC , MH , MS , MO , MN and MAC = The mass percentage of C , H , S,


charcoal were placed in a small alumina crucible and weight was re-
corded. The initial oven temperature was set as 25 °C and final tem- O, N and AC , respectively.
perature was at 1000 °C. The heating rate was 10°/min under 30 ml/
9H MC
min nitrogen (N2) purging. The ignition and the peak temperature for LHVbiomass = HHVbiomass hg +
100 100 (2)
both samples were detected by the first derivative of TGA curve called
DTG curve.
The elemental analysis was carried out by determining the ratio of
Where , LHVbiomass = The lower heating value of biomass ( );
MJ
kg

elements within the feedstocks through CHNS elemental analyzer H and MC = The hydrogen and moisture content of biomass,
(EQPCL 200 Elementar, Germany, Vario Macro Cube). The oxygen (O2) respectively (%).
was estimated by the subtraction method. During analysis, dried
MJ
hg = The latent heat of steam in the same unit as HHV ; i . e. 2.260 kg
.

336
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Fig. 1. Experimental setup of downdraft reactor


modified after [13,25]: (a) schematic diagram of a
pilot scale downdraft reactor: (i) coconut shell and
charcoal (ii) air blower, (iii) rotameter, (iv) thermo-
couples, (v) downdraft reactor, (vi) gas flare points,
(vii) cyclone separator, (viii) cooling heat exchanger,
(ix) filter, (x) clean gas sampling point, (xi) gas sam-
pling bag, (xii) temperature data logger, (xiii) com-
puter (b) pilot scale downdraft reactor during co-ga-
sification.

LHVgas = (H2 × 25.7 + CO × 30 + CH4 × 85.4) × 0.0042 (3) gasifying agent of atmospheric air was contacted with the thermal
disintegrated feedstocks, and heat released through oxidation reactions
Where, LHVgas = The lower heating value of produced syngas ( );
MJ
kg in the temperature ranges from 700 °C to 1500 °C [32]. In the reduction
H2, CO and CH4 = The hydrogen, carbon monoxide and methane , zone with oxygen starve condition, final reactions occurred between
hot gases and char. In this zone, hot gases and char were converted into
respectively (mol %).
stored chemical energy at the temperature ranges from 800 °C to
1000 °C [19,32].
HHVgas = (H2 × 30.52 + CO × 30.18 + CH4 × 95) × 0.0041868 (4)

Where, HHVgas = The higher heating value of produced syngas ( );


MJ
kg 3. Results and discussion
H2, CO and CH4 = The hydrogen, carbon monoxide and methane ,
3.1. Feedstocks characterization
respectively (mole %).
In this work, potential feedstock for co-gasification of coconut shell
2.2. Experimental setup (lignocellulosic biomass) and byproduct charcoal were characterized
for bioenergy potentiality. The proximate and ultimate characterization
The experimental setup of a pilot scale downdraft reactor system results are shown in Table 2. Fig. 2a–b shows the TGA and DTG curve
(50 kW) is shown in Fig. 1. The design parameters of reactor has been for coconut shell and charcoal. From the analytical result, it is shown
taken from the previous study that focused on the co-gasification of that lignocellulosic biomass of coconut shell has relatively low MC and
empty fruit bunch using a downdraft reactor [25]. In this study, lig- AC than charcoal (Table 2, Fig. 2), which consistent with the potential
nocellulosic biomass of coconut shell was used as the key feedstock. The usage for bioenergy production through gasification [33]. VM and FC
charcoal is another feedstock that treats as a catalyst which enhance the are the most important parameters for bioenergy production which
gasification process. The reactor body is made up of a mild steel (MS). signify fuel quality. VM is the main sources of tar compounds (Fig. 7b,
The used gasifying agent was atmospheric air (O2). The height and Table 3 and Figs. S3-S4, supplementary data) and charcoal (or the fix
diameter of the reactor was 1000 and 400 mm, respectively, and its carbon) is the sources of clean syngas. In this analysis, both feedstocks
throat angle was around 70°. The atmospheric air was injected by an have more than 50% of VM. Loh [34] reported that higher VM con-
electric blower into the reactor through the oxidation zone. The reactor taining biomasses are more potential for syngas production. The FC of
temperature was recorded by K-type thermocouples that was connected coconut shell (37.80%) is more than the charcoal (20.60%). Therefore,
with drying, pyrolysis, oxidation and reduction zone. The air flow rate FC containing charcoal might be the potential mixture with coconut
was monitored by a rotameter attached to the inlet pipe, and the shell for co-gasification with coconut shell and it also increases the
maximum flow rate capacity was upto 60 m3/h. The cyclone separator combustion time by releasing CO, H2 and CH4.
was used for the separation of solid particles from the produced syngas. The ultimate analytical results of both feedstocks are presented in
Produced hot syngas was cooled using a heat exchanger (Fig. 1). Ac- Table 2. This analysis assisted to know the theoretical amount of gas
cording to the previous study, the reactor included four individual reactants that are essential for complete fuel conversion. Analytical
zones of drying, pyrolysis, oxidation and reduction from top to bottom, results stated that coconut shell contained less than 50% (43.90%) of
respectively, and various type of complex chemical reactions (Table 1, carbon (C) and for charcoal it was more than 50% (55.43%). When
Eqs. (5)–(10)) were occurred during gasification. biomass contacted with oxygen (air) inside the downdraft reactor, then
Eq. (5) represents the overall biomass gasification reaction and in- C converted into CO and CO2 by reacting with O2 reported by Cai et al.
volves some important gasification reactions (Table 1, Eqs. (5)–(10)) [6]. Moreover, the coconut shell has higher H percentage (4.63%) than
were followed in this co-gasification study. Moisture content was eva- charcoal (1.05%). Consequently, coconut shell is the main source of H
porated from feedstocks in the drying zone, and the temperature in this that is converted into H2 and H2O. These two components of H2 and CO
zone was 100 °C–200 °C [32] and partially vapour was converted to are the key component for syngas production from any materials that
hydrogen inside the reactor. Subsequently, in the pyrolysis zone, vo- are formed from gasification at high temperature of ∼900 °C [35]. The
latile matter decreased within the temperature ranges from 150 °C to S content of both feedstocks were observed of very least amount (0.04%
700 °C [32], and thermal disintegration of the feedstocks was taken and 0.06%) that were emitted in the form of SOx. The presence of few
place and converted to volatile gases and char (Fig. 1). Thereafter, amount of sulfur in the feedstock may contribute potential usage for co-

337
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Table 1
Main chemical reactions involved in the co-gasification of coconut shell with charcoal [56,57].
Name of reaction Chemical reactions No. of equation

Overall reaction CHx Oy (coconut shell + charcoal) + O2 (atmospheric air ) = CH4 + CO + CO2 (5)

+ H2 O (untreated steam) + C (char ) + tar


Char combustion
Partial combustion 1
C + O2 = CO; H= 111KJ /mol (6)
2
Total combustion C + O2 = CO2; H= 394KJ /mol (7)
Char gasification
Boudouard equilibrium C + CO2 = 2CO; H = 173KJ /mol (8)
Water-gas reaction C + H2 O = CO + H2; H = 131KJ /mol (9)
Methanation reaction C + 2H2 = CH4; H = 75KJ /mol (10)
Homogeneous reactions
CO oxidation 1
CO + O2 = CO2; H= 283KJ /mol (11)
2
H2 oxidation H2 +
1
O2 = H2 O; H= 283KJ /mol (12)
2
CH4 oxidation CH4 + 2O2 = CO2 + 2H2 O; H = 206KJ /mol (13)
Water-gas shift reaction CO + H2 O = CO2 + H2; H = 41KJ /mol (14)
Methanation CO + 3H2 = CH4 + H2 O; H = 41KJ / mol (15)

curve (Fig. 2a–b). The TGA curve for charcoal shows the gentle mass
Table 2 loss of these materials and it was steady state till 800 °C. Subsequently,
Proximate and ultimate analysis of coconut shell and charcoal with their lignin part of the feedstock was decomposed within the temperature
heating value calculation. ranges from 350 °C to 1000 °C. The significant peak in the charcoal
Analysis Coconut shell Charcoal [25] shows at the temperature of 900 °C. It is indicated that for increasing
the ratios of charcoal with coconut shell in the co-gasification process
Proximate Analysis (wt.%) the minimum required temperature should be 1000 °C for the complete
Moisture content 6.50 7.00 conversion of feedstock.
Volatile Matter 50.70 67.40
Fixed Carbonb 37.80 20.60
Fig. 2c represents the van Krevelen diagram which represents the
Ash content 5.00 6.00 feedstock at atomic ratio of H/C and O/C based on ultimate analytical
Ultimate Analysis (wt.%) results. This analysis reveals the fuel elemental composition as regards
Carbon 43.90 55.43 the other three key components. In the diagram the main classes of
Hydrogen 4.63 1.05
solid fuels with studied feedstock samples are shown. This diagram also
Nitrogen 1.29 1.03
Sulphur 0.04 0.06 indicated that the higher content of H and O in the feedstocks are re-
Oxygenb 50.15 42.42 sponsible for higher VM. Therefore, it can be concluded some general
Heating value (MJ/kg) trends: the heating value of the selected feedstocks increases with the
Higher Heating Value (HHV) 15.47 16.07 contents of H and C, while the fuel reactivity increases with the con-
Lower Heating Value (LHV) 14.29 15.70
tents of H and O.
b= by difference. Fig. 2d represents temperature profile with time during coconut
shell co-gasification with charcoal. In the drying zone, temperature
ranges from 29 to 270 °C, where evaporation take place and vapour
gasification that would be environmentally friendly [34]. Although the converted into hydrogen during gasification [39]. In the pyrolysis zone,
presence of nitrogen (1.29% and 1.03%) and higher oxygen content thermal disintegration of coconut shell and charcoal mixtures were
(50.15% and 42.42%) decreases the energy content of feedstock due to occurred and that were converted into volatile gases and char. When
their insignificant support during combustion [36]. atmospheric air (O2) come in contact with the feedstocks in the oxi-
The higher and lower heating value of coconut shell and charcoal dation zone, various compounds react with each other that were oc-
were estimated based on Eq. (1) and Eq. (2) shown in Table 2. The curred inside the reactor (Table 1). The temperature in this zone was
higher heating value of charcoal (16.07 MJ/kg) is higher than that of around 900 °C under starved oxygen conditions. The final reactions
coconut shell (15.47 MJ/kg), and this calorific value is lower than that were occurred in the reduction zone in an inadequate of oxygen con-
of conventional fossil fuels (i.e. coal) is about 30 MJ/kg [37]. Therefore, dition, leading to reactions of hot gases with char. As a result, hot gases
these two potential feedstocks are very important for the conversion of and char is changed into the chemical energy in the form of syngas and
high energy fuel through thermochemical process such as co-gasifica- the temperature of produced gases were decreased gradually [19,39].
tion. In this study, the proximate and ultimate analytical results are In the diagram (Fig. 2d) it was clearly revealed that the temperature
good agreement with other biomass characterization results [38]. increasing with time and after the complete conversion of all feed-
Consequently, charcoal would be the potential cause for syngas pro- stocks, temperature decreases gradually due to the cooling of produced
duction if it is blended with coconut shell with suitable ratios during the syngas.
co-gasification process. The characterization results of coconut shell and charcoal using
The DTG curve (Fig. 2a–b) of coconut shell and charcoal with the TEM, FTIR and XRD analyses are shown in Fig. 3. TEM images indicated
corresponding mass loss of 6.5 (%) and 7 (%), respectively as indicated that the existence of the aggregates of nanoparticles and cellulose/
by the TGA, refers to the moisture content. After the loss of moisture hemicellulose in the coconut shell and charcoal (Fig. 3b–c and
content, the remaining materials in coconut shell and charcoal to be Fig. 3e–f). The diameter of the nanoparticles was 10–25 nm (avg.), si-
thermally stable up to a temperature close to 200 °C and 300 °C, re- milar results were reported in literature [40]. Fig. 3g represents the
spectively. Thereafter, the rate of pyrolysis increased until it reaches a representative FTIR spectra of raw feedstocks: coconut shell and char-
maximum rate of about 0.5 mass % per °C at 300 °C at coconut shell and coal. The adsorption bands at 3437.09 cm−1 and 3455.04 cm−1 ob-
in charcoal it was 400 °C. In this stage, cellulose and hemi-cellulosic served in the coconut shell and charcoal, respectively which were im-
materials were started to decompose as shown in the TGA and DTG posed to the OeH stretching reported by Lin et al. [41]. The coconut

338
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Fig. 2. Proximate and ultimate analytical results-based analysis: (a) Thermogravimetric (TGA) and differential thermogravimetric (DTG) curve of coconut shell (b)
TGA and DTG curve of charcoal (c) Van Krevelen diagram at the atomic ratio of H/C and O/C for coconut shell and charcoal, respectively (d) temperature profile of
various downdraft reactor zones.

Table 3 insignificant band at 1077.03 cm−1 in charcoal, which indicated CH2


Main chemical compounds of co-product tar produced from co-gasification of symmetric stretching [42]. Another insignificant band showed at
coconut shell and charcoal. 608.01 cm−1 is associated to CeOeC stretching [42] detected in co-
Peak No R.T. (min) Compound name % Area conut shell, and absent in charcoal. Moreover, coconut shell containing
bands that are associated with cellulose, hemicellulose or lignin, similar
1 4.710 Butyrolactone 1.53 trends followed by Liu et al. [42], where as insignificant bands re-
2 6.606 Phenol 57.20
presented by charcoal. Ultimately, coconut shell is most likely consisted
3 8.375 2-Ethylbutylamine 2.42
4 9.635 2-Chloro-5,5-dimethyl-1-phenyl-3-h exen-1-ol 2.99 of CeH, OeH and CH2 stretching, while the charcoal presented only OH
5 10.824 Phenol, 3-methyl- 5.42 [25,43]. Fig. 3h represents the XRD patterns for both feedstocks. The
6 11.441 Phenol, 2-methoxy- 4.61 peak positions at 29.48, 44.11, 64.42, 65.55 and 77.49 for the coconut
7 12.039 Acetaldehyde 1.05
shell and the appearance of peak position at 26.66, 28.58, 30.21, 31.48,
8 15.615 1,1-Difluoro-2-methyl-2-methoxycyc lopropane 0.85
9 16.970 1-Hexen-3-yne, 2,5,5-trimethyl- 1.58
33.96, 35.48, 42.05, 50.31 and 67.98 in the XRD of charcoal shows the
10 18.185 Naphthalene 1.20 presence of nanoparticles (i.e. carbon).
11 18.707 Creosol 2.41 Fig. 4a–b exhibits the XPS spectrum of coconut shell and charcoal.
12 19.305 Catechol 6.17 Significant peak of C, O, N and S were observed on coconut shell surface
13 19.502 Catechol 4.02
(Fig. 4ai-5aiv), whereas C and O were identified on the charcoal sur-
14 24.477 5-Acetyl-2-furanmethanol 1.29
15 25.069 1,3,5-Cyclooctatriene 1.58 faces (4bi-4bii). In this analysis, C1s and O1s curves were fitted both on
16 26.443 Phenol, 4-ethyl-2-methoxy- 2.10 both feedstocks that indicate an asymmetric C1s at high binding energy
17 26.799 1,3,5,7-Cyclooctatetraene 0.34 signifying the exposure of oxygen-containing groups, similar trends
18 28.192 1,2-Benzenediol, 4-methyl- 2.70 followed in the previous study by Monir et al. [25]. The C1s peak in
19 29.268 2-Pentanone 0.54
coconut shell sample was deconvoluted into three different peaks,
where binding energies were 292.99 eV, 294.44 eV and 296.65 eV
shell presented the supplementary bands at 2922.33 cm−1 attributed to (Fig. 4ai). Four different peaks in coconut shell at the binding energy of
CeH stretching [25], which was absent in charcoal. In the FTIR spectra, 536.02, 537.94, 540.38 eV and 540.90 eV were indicated to the O1s
it is noted that bands at 1734.28 cm−1 and 1638.04 cm−1 in coconut peaks [25,44]. The oxygenated functional groups were also agreed with
shell and comparatively insignificant band at 1635.91 cm−1 agrees to the outcomes of proximate, ultimate and FTIR analytical results. The
the stretching vibration of the bending mode of OeH in charcoal stu- XPS narrow scans for N1s and S2p for coconut shell are given in
died by Liu et al. [42]. In addition, two significant band at Fig. 4aiii-iv. The spectrum of coconut shell represents a peak at
1252.74 cm−1 and 1046.29 cm−1 were also observed in coconut shell, 398.44 eV, 399.21 eV, 399.71 eV, 400.22 eV, 401.49 eV and 402.43 eV

339
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Fig. 3. TEM, FTIR and XRD analysis of coconut shell and charcoal: (a) raw feedstock of coconut shell (b,c) TEM images of coconut shell (d) raw feedstock of charcoal
(e,f) TEM images of charcoal (g) FTIR curve for coconut shell and charcoal (h) XRD curve for coconut shell and charcoal.

corresponding to N1s species typically exhibits in high-surface samples. 3.2. Syngas production
The signal at 160.89 eV, 161.97 eV, 162.39 eV, 162.96 eV, 163.47 eV
and 164.27 eV binding energy, which is ascribed to S2p species in the The experiments were run with the mixture of coconut shell and
sample (Fig. 4aiv). These curves agree well with the literature value charcoal with the blending ratio of 100:0, 90:10, 80:20 and 70:30. The
[25,45]. In the previous study, deconvolution of the C1s peak of char- experimental results are represented in Fig. 5a–c. The produced syngas
coal corresponds to the different carbon-based functional groups as consists of combustible (H2, CO, CH4) and non-combustible (CO2, N2)
shown Fig. 4b that supported the existence of CeC/C]C, CeOeC, re- compounds analyzed by GC-TCD (Fig. S1, supplementary data). The
spectively (Fig. 4bi) and this result is agreed with the literature [46]. mole percentage of syngas (H2, CO, CO2, CH4) are shown in Fig. 5a. In
From elemental analysis (Table 2) it was evident that total carbon Fig. 5b represents the yield composition of H2 and CO. The syngas ratio
content of coconut shell and charcoal is 43.90% and 55.43% respec- (H2 and CO) for the various mixture of coconut shell and charcoal of
tively in the form of condensed aromatic rings and aliphatic groups. 100:0, 90:10, 80:20, and 70:30 are 1.08, 1.09, 1.10 and 1.10, respec-
Fig. 4bii represents three different peaks in the charcoal with the tively.
binding energy of 542.79 eV, 544.32 eV and 546.19 eV were assigned to The mole percentage of H2 and CO were increased gradually by
the O1s presence in the charcoal sample, similar studied were done in increasing the mixture of charcoal with coconut shell. With the increase
the previous work [25]. of charcoal from 0 to 30%, concentration of H2 and CO increased

340
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Fig. 4. X-ray photoelectron spectroscopy (XPS) curve-fitted for coconut shell and charcoal: (a) wide scan of coconut shell (ai) narrow scan of C1s (aii) narrow scan of
O1s (aiii) narrow scan of N1s (aiv) narrow scan of S2p (b) wide scan of charcoal (bi) narrow scan of C1s (bii) narrow scan of O1s.

7.07–8.44% and 13.04–15.38%, respectively. In contrary, the con- concentration insignificantly occurred in the producer gas [48].
centration of CO2 decreased with increasing the charcoal ratio with The exergy analysis was performed for design and analysis of bio-
coconut shell. Both feedstocks contained alkane as evident from FTIR mass conversion. The exergetic efficiency is the ratio of total exergy
analysis (Fig. 3g). Further it was confirmed by GC-MS analysis (Fig. S2, produced from the reactor and total exergy input to the reactor as
Supplementary data) which was consistent of FTIR result. Nevertheless, follows in Eq. (5) [18]:
there are no remarkable effects on CH4 concentration of various co-
conut shell and charcoal ratios. The similar trend followed for CH4 Ex prod
ex =
concentration fairly consistent with the literature [47]. The increasing Ex feedstocks + Exagent (5)
trend for H2 and CO concentration within the reactor was occurred due
to the change of temperature both oxidation and reduction zone. It Where, Ex prod and Ex feedstocks
occurred because of synergistic effect between coconut shell = The exergies of produced syngas and inlet feedstocks , respectively ,
and charcoal. As a result of limited methanation reactions CH4
while Exagent is the exergy of gasifying agent (i . e. air ).

341
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

Fig. 5. Product analysis produced from the co-gasification of coconut shell and charcoal: (a) syngas composition with various ratios of charcoal with coconut shell (b)
H2 and CO yield (%) for produced syngas (c) HHV and LHV for produced syngas (d) exergy efficiency for co-gasification of coconut shell and charcoal with the ratios
of 100:0, 90:10, 80:20 and 70:30.

As the co-gasifying agent was atmospheric air and it was inserted C > 0.667 were estimated by the following equation (Eq. (9)) [18].
into the reactor through oxidation zone at the atmospheric condition,
gasifying agent was neglected for the calculation of exergy efficiency fm = (LHVfm + 2442 × Wfm) (9)
reported by Prins and Ptasinski [49]. The exergy of syngas (Exprod) is
Where , LHVfm = The lower heating value of fuel mixture (coconut shell and charcoal );
the summation of potential, kinetic, physical and chemical exergy. In
Wfm = The mass fraction of moisture in the fuel mixture
this analysis, potential and kinetic exergy covers independent tem-
perature and pressure. Consequently, their role to exergy were insig- (coconut shell and charcoal).
nificant. Therefore, the physical and chemical exergy attributed the In this analysis, Eq. (10) was used to calculate β value by comparing
major contributors that were considered for the estimation of exergy as the mass fraction of H, C, O and N in the feedstocks of coconut shell and
shown in Eq. (6). charcoal.

Ex = Ex phy + Exchem (6) 1.0438 + 0.1882 C


H O
0.2509 C 1 + 0.7256 C
H
+ 0.0383 C
N

= O
Eq. (7) represents the generalized equation for the calculation of 1 0.3035 C (10)
specific physical exergy are as:
Fig. 5d represents the exergy efficiency for the co-gasification of
ph = (h h0) T0 (s s0 ) (7) coconut shell and charcoal with the ratios of 100:0, 90:10, 80:20 and
Where, h and s = The enthalpy and entropy at any temperature (T ) and pressure (P );
70:30. It is clearly shown that exergy efficiency was increasing with
h 0 and s0 = The enthalpy and entropy at dead state (T0 = 298.15 K , increasing the mixture of charcoal with coconut shell. The exergy ef-
ficiencies were varied from 35.96% to 50.85% when charcoal with
P0 = 1 atm .). coconut shell was changed from 0% to 30%. It was occurred because of
Eq. (8) shows specific chemical exergy for the estimation of an ideal the increasing nature of physical and chemical exergy in syngas. As a
gas mixture are as follows: result, the effect on reactor temperature depends on the physical exergy
of the producer gas. The exergy for various mixture of coconut shell and
= xi + RT0 x i In x i
ch . m ch . i
(8) charcoal were estimated using Eq. (9). Based on Eqs. (3) and (4), HHV
i i
and LHV for the coconut shell was 15.47 MJ/kg and 14.29 MJ/kg, re-
Where , x i = The mole fraction of ith species; R is the universal gas constant; spectively and in charcoal it was 16.07 MJ/kg and 15.58 MJ/kg, re-
spectively, similar trend was followed in the literature [25]. It is ob-
KJ
ch . i = The standard chemical exergy of ithspecies in kmol .
The exergy of solid biomass with the mass ratio of 2.67 > O/ served that the value of LHV for coconut shell is little bit lower than

342
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

like sodium, sulfur, calcium are in trace amount (Fig. 6). Fig. 7a shows
the FTIR spectra of tar sample obtained from co-gasification process.
The spectral band at 3351.02 cm−1 represents OH stretching. The
bands located in 2925.66 cm−1 and 2854.37 cm−1 that are corre-
sponding to the distribution of aliphatic CH2 groups. The bands at
1361.33 cm−1 and 1361.33 cm−1 matched to the distribution of ali-
phatic CH3 groups. In addition, C]C ring stretching bands at
1606.23 cm−1 and 1514.52 cm−1 were observed in tar sample. In the
presence of phenol, it is known that this band is strongly exalted. These
bands are consistent with the literature value reported by Michel et al.
[52]. In addition, the infrared spectra show the presence of phenolic
compounds (Fig. 7a).
Fig. 6. SEM image with EDX of the co-product tar produced from co-gasifica- Fig. 7b represents various peaks of organic compounds of co-pro-
tion of coconut shell and charcoal in the downdraft reactor.
duct tar produced from the co-gasification of coconut shell and char-
coal. Table 3 represents the tentative list of tar compounds that are
charcoal. Nevertheless, β value was decreasing when the charcoal with identified by GC-MS analysis. The main compounds are numbered in
coconut shell ratios were increasing. Therefore, the value of LHV, MC Fig. 7b and each compound has the same number shown in Table 3. It
and β were found to be proportional to the overall exergy. can be seen that the major compounds are phenol cover maximum peak
area (57.20%). The other significant peak observed in Fig. 7b at the
3.3. Co-product analysis retention time of 10.824 (min), 11.441 (min), 19.305 (min), 19.502
(min) which denotes the corresponding compounds of Phenol, 3-me-
Tar is the co-product produced from the co-gasification of coconut thyl-, 2-methoxy- and Catechol (Table 3). In addition, other significant
shell and charcoal. It is the mixture of various condensable hydro- compounds were also observed in the tar samples are as Butyrolactone,
carbons, including aromatic compounds with up to five rings as well as 2-Ethylbutylamine, 2-Chloro-5,5-dimethyl-1-phenyl-3-h exen-1-ol,
Polycyclic Aromatic Hydrocarbons (PAHs) having low or no water so- Acetaldehyde, 1-Hexen-3-yne, 2,5,5-trimethyl-, 5-Acetyl-2-fur-
lubility characteristics (hydrophobic substances) [50]. Generally, the anmethanol, 1,3,5-Cyclooctatriene, Phenol, 4-ethyl-2-methoxy-, 1,2-
presence of tar affects the quality of syngas [50,51]. The concentration Benzenediol, 4-methyl-. Some other insignificant peaks were also
of tar in syngas is depended on the type of feedstock, moisture content identified (Fig. 7b). Similar results are found in the literature
and reactor design. [50,51,53].
Fig. 6 represents the surface morphology and elemental composition The normalized integration spectra of 13C NMR and 1H NMR for the
of tar sample. The SEM image of tar shows a smooth surface with minor co-gasification-based co-product tar are shown in Table 4 and Figs. S3-
pitting due to the release of gas under pressure during gasification. The S4 (supplementary data). In the 13C NMR spectra analysis of tar shows
identified elements in tar samples are carbon (81.76%), oxygen that 100–157 ppm (aromatic region), is represented various types of
(17.89%), aluminum (0.17%), potassium (0.10%) and other elements carbon to eOH in phenolic compounds. Comparable spectra are also
found in the literature value [52,54]. Seeing the spectra of aromatic
hydrogen from 1H NMR (Table 4 and Fig. S4, supplementary data), it
can be shown that the co-product tars are rich in aromatic hydrogen.
The determination of tar compounds by GC-MS analysis has allowed
this study to identify the main peaks in 13C and 1H NMR analysis. Due
to the effect of little concentration on chemical shifts makes the
quantitative analysis of the spectra difficult. However, the several sig-
nals can be assigned and are related to the main peaks identified by GC-
MS analysis: Phenols, Butyrolactone, 2-Chloro-5, 5-dimethyl-1-phenyl-
3-h exen-1-ol, 2-Ethylbutylamine and Naphthalene. Undeniably, both
13
C NMR and 1H NMR analyses were agreed the results obtained in GC-
MS.

Table 4
13
C NMR and 1H NMR data of co-product tar compounds produced from co-
gasification of coconut shell and charcoal.
13 1
Peaks C NMR (δC ppm) H NMR (δH ppm)

1 55.695 0.658
2 115.203 1.069
3 118.360 2.720
4 119.199 3.607
5 119.788 5.037
6 129.334 5.071
7 131.798 6.568
8 137.741 6.583
9 148.664 6.598
10 155.079 6.613
11 157.464 6.704
12 166.676 6.958
13 7.683
Fig. 7. (a) FTIR analysis for the identification of functional group of co-product 14 7.700
15 8.147
tar (b) GC-MS chromatogram of co-product tar produced from co-gasification of
16 8.495
coconut shell and charcoal.

343
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

From the above analysis, various types of organic compounds were produced from rice husk gasification in an entrained flow reactor, J. Clean. Prod. 95
found in the co-product tar. Identified main components were phenols, (2015) 273–280.
[10] S.K. Sansaniwal, K. Pal, M.A. Rosen, S.K. Tyagi, Recent advances in the develop-
Butyrolactone, 2-Chloro-5, 5-dimethyl-1-phenyl-3-h exen-1-ol, 2- ment of biomass gasification technology: a comprehensive review, Renew. Sustain.
Ethylbutylamine and Naphthalene as shown in Tables 3 and 4. A con- Energy Rev. 72 (2017) 363–384.
clusion has been reached by several researchers [35,51,53,55] that the [11] A. Chaurasia, Modeling of downdraft gasification process: studies on particle geo-
metries in thermally thick regime, Energy 142 (2018) 991–1009.
water phase exists in oxygenated organic components separate during [12] V.S. Sikarwar, M. Zhao, P. Clough, J. Yao, X. Zhong, M.Z. Memon, N. Shah,
the gasification process. Thus, it can be suggested that the water phase E.J. Anthony, P.S. Fennell, An overview of advances in biomass gasification, Energy
should remove from the component to produce clean and sustainable Environ. Sci. 9 (10) (2016) 2939–2977.
[13] M.U. Monir, A. Yousuf, A.A. Aziz, S.M. Atnaw, Enhancing Co-gasification of coconut
energy. shell by reusing char, Indian J. Sci. Technol. 10 (6) (2017) 1–4.
[14] D. Mallick, P. Mahanta, V.S. Moholkar, Synergistic Effects in Gasification of Coal/
4. Conclusion Biomass Blends: Analysis and Review, Coal and Biomass Gasification, Springer,
2018, pp. 473–497.
[15] I. Adeyemi, I. Janajreh, T. Arink, C. Ghenai, Gasification behavior of coal and
Gasification of the lignocellulosic biomass of coconut shell with the woody biomass: validation and parametrical study, Appl. Energy 185 (2017)
various blending ratios of charcoal was investigated in a 50 kW 1007–1018.
downdraft reactor for the production of syngas. The temperature in the [16] M.M. Farid, M.S. Kang, J. Hwang, The effect of CO on coal–biomass co-gasification
with CO 2, Fuel 188 (2017) 98–101.
oxidation and reduction zone were increased gradually with increasing [17] D. Mallick, P. Mahanta, V.S. Moholkar, Co-gasification of coal and biomass blends:
time and it decreased with the reduction of feedstocks. The main chemistry and engineering, Fuel 204 (Supplement C) (2017) 106–128.
components of producer syngas (H2 and CO) increased its concentration [18] V.R. Patel, D. Patel, N. Varia, R.N. Patel, Co-gasification of lignite and waste wood
in a pilot-scale (10 kWe) downdraft gasifier, Energy 119 (2017) 834–844.
with increasing charcoal ratio with biomass. The CH4 concentration [19] M. Sharma, S. Attanoor, S. Dasappa, Investigation into co-gasifying Indian coal and
was relatively unchanged with increasing charcoal ratio to biomass. biomass in a down draft gasifier—experiments and analysis, Fuel Process. Technol.
The remarkable syngas ratio was observed as 1.10 (H2:CO) when the 138 (2015) 435–444.
[20] M.M.U. Monir, H.Z. Hossain, Coal mine accidents in Bangladesh: its causes and
coconut shell and charcoal ratio was 70:30. Gas with an LHV of remedial measures, Int. J. Econ. Environ. Geol. 3 (2012) 33–40.
2.81–3.27 MJ/Nm3 was observed at a reduction zone temperature of [21] R.C. Milici, R.M. Flores, G.D. Stricker, Coal resources, reserves and peak coal pro-
788 °C. The overall exergy efficiencies were higher with higher mixture duction in the United States, Int. J. Coal Geol. 113 (2013) 109–115.
[22] F.-X. Collard, J. Blin, A review on pyrolysis of biomass constituents: mechanisms
of charcoal to the lignocellulosic biomass of coconut shell. The co- and composition of the products obtained from the conversion of cellulose, hemi-
product of tar was also characterized and identified components mainly celluloses and lignin, Renew. Sustain. Energy Rev. 38 (2014) 594–608.
are phenols, naphthalene, and catechol. Therefore, coconut shell and [23] A. Adrados, A. Lopez-Urionabarrenechea, E. Acha, J. Solar, B.M. Caballero, I. de
Marco, Hydrogen rich reducing gases generation in the production of charcoal from
charcoal could be the potential source of renewable bioenergy for the
woody biomass carbonization, Energy Convers. Manag. 148 (Supplement C) (2017)
fulfilment of future energy demand. 352–359.
[24] S.M. Atnaw, S.A. Sulaiman, S. Yusup, Biomass Gasification, Waste Biomass
Acknowledgments Management–a Holistic Approach, Springer, 2017, pp. 159–185.
[25] M.U. Monir, A. Abd Aziz, R.A. Kristanti, A. Yousuf, Co-gasification of empty fruit
bunch in a downdraft reactor: a pilot scale approach, Bioresour. Technol. Rep. 1
The authors would like to thanks to the Faculty of Engineering and (2018) 39–49.
Technology (FTeK), Universiti Malaysia Pahang, Malaysia for providing [26] B. Cantero-Tubilla, D.A. Cantero, C.M. Martinez, J.W. Tester, L.P. Walker,
R. Posmanik, Characterization of the solid products from hydrothermal liquefaction
lab (Energy Management Lab) facilities. Financial support (RDU Grant of waste feedstocks from food and agricultural industries, J. Supercrit. Fluids 133
No. RDU160317 and GRS Grant No. PGRS170370) received from the (Part 2) (2017).
Faculty of Engineering Technology (FTek), Universiti Malaysia Pahang [27] A.S. Baharuddin, A. Sulaiman, D.H. Kim, M.N. Mokhtar, M.A. Hassan, M. Wakisaka,
Y. Shirai, H. Nishida, Selective component degradation of oil palm empty fruit
(UMP), Malaysia, for completing this research is greatly acknowledged. bunches (OPEFB) using high-pressure steam, Biomass Bioenergy 55 (Supplement C)
(2013) 268–275.
Appendix A. Supplementary data [28] S.A. Channiwala, P.P. Parikh, A unified correlation for estimating HHV of solid,
liquid and gaseous fuels, Fuel 81 (8) (2002) 1051–1063.
[29] P. Basu, Biomass Gasification and Pyrolysis: Practical Design and Theory, Academic
Supplementary data to this article can be found online at https:// press, 2010.
doi.org/10.1016/j.biombioe.2018.10.006. [30] M. Shahbaz, S. Yusup, D.A. Inayat, M. Ammar, D.O. Patrick, A. Pratama, S.R. Naqvi,
Syngas Production from Steam Gasification of Palm Kernel Shell with Subsequent
CO2 Capturing Using CaO Sorbent: an Aspen Plus Modelling, Energy & Fuels, 2017.
References [31] C.F. Valdés, G. Marrugo, F. Chejne, J.I. Montoya, C.A. Gómez, Pilot-scale fluidized-
bed co-gasification of palm kernel shell with sub-bituminous coal, Energy Fuel. 29
[1] M. Paiman, N. Hamzah, S. Idris, N. Rahman, K. Ismail, Synergistic effect of Co- (9) (2015) 5894–5901.
utilization of coal and biomass char: an overview, IOP Conference Series: Materials [32] X. Zhang, H. Li, L. Liu, C. Bai, S. Wang, J. Zeng, X. Liu, N. Li, G. Zhang,
Science and Engineering, IOP Publishing, 2018, p. 012003. Thermodynamic and economic analysis of biomass partial gasification process,
[2] R. Kurniawan, S. Managi, Coal consumption, urbanization, and trade openness Appl. Therm. Eng. 129 (2018) 410–420.
linkage in Indonesia, Energy Pol. 121 (2018) 576–583. [33] H. Yang, H. Chen, 11 - biomass gasification for synthetic liquid fuel production A2 -
[3] Y. Li, X. Shi, B. Su, Economic, social and environmental impacts of fuel subsidies: a Luque, Rafael, in: J.G. Speight (Ed.), Gasification for Synthetic Fuel Production,
revisit of Malaysia, Energy Pol. 110 (2017) 51–61. Woodhead Publishing, 2015, pp. 241–275.
[4] K. Kundu, A. Chatterjee, T. Bhattacharyya, M. Roy, A. Kaur, Thermochemical [34] S.K. Loh, The potential of the Malaysian oil palm biomass as a renewable energy
conversion of biomass to bioenergy: a review, in: A.P. Singh, R.A. Agarwal, source, Energy Convers. Manag. 141 (2017) 285–298.
A.K. Agarwal, A. Dhar, M.K. Shukla (Eds.), Prospects of Alternative Transportation [35] A.S. Al-Rahbi, P.T. Williams, Hydrogen-rich syngas production and tar removal
Fuels, Springer Singapore, Singapore, 2018, pp. 235–268. from biomass gasification using sacrificial tyre pyrolysis char, Appl. Energy 190
[5] F.H. Isikgor, C.R. Becer, Lignocellulosic biomass: a sustainable platform for the (2017) 501–509.
production of bio-based chemicals and polymers, Polym. Chem. 6 (25) (2015) [36] S. Czernik, A. Bridgwater, Overview of applications of biomass fast pyrolysis oil,
4497–4559. Energy Fuel. 18 (2) (2004) 590–598.
[6] J. Cai, Y. He, X. Yu, S.W. Banks, Y. Yang, X. Zhang, Y. Yu, R. Liu, A.V. Bridgwater, [37] S. Channiwala, P. Parikh, A unified correlation for estimating HHV of solid, liquid
Review of physicochemical properties and analytical characterization of lig- and gaseous fuels, Fuel 81 (8) (2002) 1051–1063.
nocellulosic biomass, Renew. Sustain. Energy Rev. 76 (2017) 309–322. [38] A.J. Tsamba, W. Yang, W. Blasiak, Pyrolysis characteristics and global kinetics of
[7] T. Bhaskar, B. Bhavya, R. Singh, D.V. Naik, A. Kumar, H.B. Goyal, Chapter 3 - coconut and cashew nut shells, Fuel Process. Technol. 87 (6) (2006) 523–530.
thermochemical conversion of biomass to biofuels A2 - Pandey, Ashok, in: [39] L. Wei, Experimental study on the effects of operational parameters of a downdraft
C. Larroche, S.C. Ricke, C.-G. Dussap, E. Gnansounou (Eds.), Biofuels, Academic gasifier, Masters Abstracts Int. 47 (01) (2005).
Press, Amsterdam, 2011, pp. 51–77. [40] E. Longhin, M. Gualtieri, L. Capasso, R. Bengalli, S. Mollerup, J.A. Holme,
[8] W.-C. Yan, Y. Shen, S. You, S.H. Sim, Z.-H. Luo, Y.W. Tong, C.-H. Wang, Model- J. Øvrevik, S. Casadei, C. Di Benedetto, P. Parenti, M. Camatini, Physico-chemical
based downdraft biomass gasifier operation and design for synthetic gas produc- properties and biological effects of diesel and biomass particles, Environ. Pollut.
tion, J. Clean. Prod. 178 (2018) 476–493. 215 (2016) 366–375.
[9] Y. Zhang, Y. Zhao, X. Gao, B. Li, J. Huang, Energy and exergy analyses of syngas [41] G. Lin, H. Yang, X. Wang, Y. Mei, P. Li, J. Shao, H. Chen, The moisture sorption
characteristics and modelling of agricultural biomass, Biosyst. Eng. 150 (2016)

344
M.U. Monir et al. Biomass and Bioenergy 119 (2018) 335–345

191–200. Reduction of tar generated during biomass gasification: a review, Biomass


[42] C.-F. Liu, J.-L. Ren, F. Xu, J.-J. Liu, J.-X. Sun, R.-C. Sun, Isolation and character- Bioenergy 108 (2018) 345–370.
ization of cellulose obtained from ultrasonic irradiated sugarcane bagasse, J. Agric. [51] V. Pallozzi, A. Di Carlo, E. Bocci, M. Carlini, Combined gas conditioning and
Food Chem. 54 (16) (2006) 5742–5748. cleaning for reduction of tars in biomass gasification, Biomass Bioenergy 109
[43] A. Meng, H. Zhou, L. Qin, Y. Zhang, Q. Li, Quantitative and kinetic TG-FTIR in- (2018) 85–90.
vestigation on three kinds of biomass pyrolysis, J. Anal. Appl. Pyrol. 104 (2013) [52] R. Michel, S. Rapagnà, P. Burg, G. Mazziotti di Celso, C. Courson, T. Zimny,
28–37. R. Gruber, Steam gasification of Miscanthus X Giganteus with olivine as catalyst
[44] D. Wang, Z. Geng, B. Li, C. Zhang, High performance electrode materials for electric production of syngas and analysis of tars (IR, NMR and GC/MS), Biomass Bioenergy
double-layer capacitors based on biomass-derived activated carbons, Electrochim. 35 (7) (2011) 2650–2658.
Acta 173 (Supplement C) (2015) 377–384. [53] W.T. Tsai, M.K. Lee, Y.M. Chang, Fast pyrolysis of rice straw, sugarcane bagasse and
[45] Y. Cai, X. Yang, T. Liang, L. Dai, L. Ma, G. Huang, W. Chen, H. Chen, H. Su, M. Xu, coconut shell in an induction-heating reactor, J. Anal. Appl. Pyrol. 76 (1) (2006)
Easy incorporation of single-walled carbon nanotubes into two-dimensional MoS2 230–237.
for high-performance hydrogen evolution, Nanotechnology 25 (46) (2014) 465401. [54] P. Wang, L. Jin, J. Liu, S. Zhu, H. Hu, Analysis of coal tar derived from pyrolysis at
[46] C.H. Chia, S.D. Joseph, A. Rawal, R. Linser, J.M. Hook, P. Munroe, Microstructural different atmospheres, Fuel 104 (2013) 14–21.
characterization of white charcoal, J. Anal. Appl. Pyrol. 109 (Supplement C) (2014) [55] N. Xu, S.S. Kim, A. Li, J.R. Grace, C.J. Lim, T. Boyd, Investigation of the influence of
215–221. tar-containing syngas from biomass gasification on dense Pd and Pd–Ru mem-
[47] K. Kumabe, T. Hanaoka, S. Fujimoto, T. Minowa, K. Sakanishi, Co-gasification of branes, Powder Technol. 290 (2016) 132–140.
woody biomass and coal with air and steam, Fuel 86 (5) (2007) 684–689. [56] G. Oh, H.W. Ra, S.M. Yoon, T.Y. Mun, M.W. Seo, J.G. Lee, S.J. Yoon, Gasification of
[48] J. Fermoso, B. Arias, M.G. Plaza, C. Pevida, F. Rubiera, J.J. Pis, F. García-Peña, coal water mixture in an entrained-flow gasifier: effect of air and oxygen mixing
P. Casero, High-pressure co-gasification of coal with biomass and petroleum coke, ratio, Appl. Therm. Eng. 129 (Supplement C) (2018) 657–664.
Fuel Process. Technol. 90 (7) (2009) 926–932. [57] Y. Richardson, M. Drobek, A. Julbe, J. Blin, F. Pinta, Chapter 8 - biomass gasifi-
[49] M.J. Prins, K.J. Ptasinski, Energy and exergy analyses of the oxidation and gasifi- cation to produce syngas A2 - Pandey, Ashok, in: T. Bhaskar, M. Stöcker,
cation of carbon, Energy 30 (7) (2005) 982–1002. R.K. Sukumaran (Eds.), Recent Advances in Thermo-chemical Conversion of
[50] M.L. Valderrama Rios, A.M. González, E.E.S. Lora, O.A. Almazán del Olmo, Biomass, Elsevier, Boston, 2015, pp. 213–250.

345

Potrebbero piacerti anche