Sunteți pe pagina 1din 39

Accepted Manuscript

Effect of particle size and temperature on gasification performance of coconut and


palm kernel shells in downdraft fixed-bed reactor

Ahmad Zubair Yahaya, Mahendra Rao Somalu, Andanastuti Muchtar, Shaharin


Anwar Sulaiman, Wan Ramli Wan Daud

PII: S0360-5442(19)30554-7

DOI: 10.1016/j.energy.2019.03.138

Reference: EGY 14974

To appear in: Energy

Received Date: 31 October 2018

Accepted Date: 24 March 2019

Please cite this article as: Ahmad Zubair Yahaya, Mahendra Rao Somalu, Andanastuti Muchtar,
Shaharin Anwar Sulaiman, Wan Ramli Wan Daud, Effect of particle size and temperature on
gasification performance of coconut and palm kernel shells in downdraft fixed-bed reactor, Energy
(2019), doi: 10.1016/j.energy.2019.03.138

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Effect of particle size and temperature on gasification performance of coconut and palm

kernel shells in downdraft fixed-bed reactor

Ahmad Zubair Yahayaa, Mahendra Rao Somalua*, Andanastuti Muchtara,b, Shaharin Anwar

Sulaimanc, Wan Ramli Wan Daud a,d

a Fuel Cell Institute, Universiti Kebangsaan Malaysia, 43600 UKM Bangi, Selangor,

Malaysia
b Centre for Materials Engineering and Smart Manufacturing (MERCU), Faculty of

Engineering and Built Environment, Universiti Kebangsaan Malaysia, 43600 UKM Bangi,

Selangor, Malaysia
c Department of Mechanical Engineering, Universiti Teknologi PETRONAS, Bandar Seri

Iskandar, 31750 Tronoh, Perak, Malaysia


d Research Centre for Sustainable Process Technology (CESPRO), Faculty of Engineering

and Built Environment, Universiti Kebangsaan Malaysia, 43600 UKM Bangi, Selangor,

Malaysia

Tel.: +603-89118522

Fax: +603-89118530

*corresponding author: mahen@ukm.edu.my

1
ACCEPTED MANUSCRIPT

Abstract

Gasification of coconut shell (CS) and palm kernel shell (PKS) is conducted in a batch type

downdraft fixed-bed reactor to evaluate the effect of particle size (1–3 mm, 4–7 mm, and 8–11

mm) and temperature (700, 800, and 900 °C) on gas composition and gasification performance.

The response surface methodology integrated variance-optimal design is used to identify the

optimum condition for gasification. Gas composition, which is measured using the biomass

particle size of 1–11 mm at 700–900 °C, are 8.20–14.6 vol% (H2), 13.0–17.4 vol% (CO), 14.7–

16.7 vol% (CO2), and 2.82–4.23 vol% (CH4) for CS and 7.01–13.3 vol% (H2), 13.3–17.8 vol%

(CO), 14.9–17.1 vol% (CO2), and 2.39–3.90 vol% (CH4) for PKS. At similar conditions, the

syngas higher heating value, dry gas yield, carbon conversion efficiency, and cold gas

efficiency are 4.01–5.39 MJ/Nm3, 1.50–1.95 Nm3/kg, 52.2–75.9 %, and 30.9–56.4 % for CS,

respectively, and 3.82–5.09 MJ/Nm3, 1.48–1.92 Nm3/kg, 59.0–81.5 %, and 33.0–57.1 % for

PKS, respectively. Results reveal that temperature has a greater role than particle size in

influencing the gasification reaction rate.

Keywords: Coconut shell; palm kernel shell; coconut shell; downdraft gasifier; biomass particle

size; gasification temperature

2
ACCEPTED MANUSCRIPT

1. Introduction

The world energy demand in 2018 rose more than twice the demand in 2017 and this trend

is proportional to the increase in carbon dioxide emissions [1]. With respect to global energy

security, renewable energy has received impressive gain to support and reduce people’s

dependence on fossil fuels, which are the main contributors to carbon dioxide emissions.

Biomass gasification is recognized as one of the most promising renewable energy

technologies for fueling gas turbines, internal combustion engines, or fuel cells for power

generation. This technology works by converting biomass to valuable products of gas (syngas),

solid (char), and liquid (tar).

Nowadays, biomass is abundantly available worldwide, and majority are derived from

agriculture crops. Coconut (Cocos nucifera L.) and oil palm trees (Elaeis guineensis) are

considered among the world’s main agricultural crops. According to LMC International Ltd.,

global crude oil production from coconut and oil palm trees are approximately 3.35 million

tonnes and 60.86 million tonnes, respectively (Oilseeds and Oils Report 2015). As a result,

over 70% of solid wastes were generated from post-processing of products such as trunks,

fronds, leaves, fruits, and shells. Among them, coconut shells (CSs) and palm kernel shells

(PKSs) have been considered potential biomass because of their easy handling and syngas

higher heating values (HHV) of 18.9–21.5 MJ/kg [3–5] and 16.14–18.9 MJ/kg [6–8],

respectively. Therefore, utilizing abundant solid wastes for energy generation through the

empowerment of gasification is valuable.

Gasification performance is influenced by many factors, including the chemical and

physical properties of biomass [9], gasifier design [10], and operational conditions, such as

temperature [11], equivalence ratio [12], and gasifying agent [13]. Currently, various

commercial gasifier designs, such as updraft, downdraft and fluidized-bed gasifier, have been

3
ACCEPTED MANUSCRIPT

developed. Of the developed gasifiers, the downdraft gasifier is the most sustainable for

combined heat and power generation because of low tar and particulate composition in the

produced syngas.

Studies on biomass gasification have been conducted to evaluate the effect of operating

parameters, such as temperature, residence time, steam to biomass ratio, particle size, and

biomass types, on gasification performance. Various particle size and temperature ranges have

been adopted to investigate their effect on the performance of fluidized bed [14–17] and updraft

fixed-bed [18–21] reactors. However, to the best of authors knowledge, only two studies have

examined the effect of biomass particle size on gasification performance in a downdraft fixed-

bed reactor. Patel et al. [22] studied the influence of six different lignites with particle sizes

ranging from 13–31 mm on gasification using downdraft gasifier. The study revealed that an

increase in particle size from 13 mm to 25 mm increases producer gas composition and gas

calorific value, as opposed to particle sizes between 25 mm and 31 mm. In addition, Inayat et

al. [23] investigated the effects of three different particle sizes of CS and wood chip ranging

from 5 mm to 50 mm using a downdraft gasifier. The study found that the small particle size

of CS and wood chip blending generated valuable gases and produced considerable cold gas

efficiency () and syngas HHV (HHVsyngas). However, the dry gas yield (Y) of the CS and wood

chip blending with small particle sizes showed a reverse trend. At present, limited information

is found in the literature regarding the effect of temperature (700–900 °C) and particle size (1–

11 mm) on gasification performance under a typical condition of downdraft fixed-bed gasifier.

Further investigation is necessary to understand these effects on gasification performance using

a downdraft reactor.

In the present work, air gasification of CS and PKS is experimentally investigated in a

fixed-bed downdraft reactor. The purpose of this study is to evaluate the effect of particle size

and temperature on gasification output (with reference to gas composition (CO, CO2, CH4, and

4
ACCEPTED MANUSCRIPT

H2), gasification performance (HHVsyngas, Y, ηc, and ), and mass balance (material and

carbon). In addition, the response surface methodology (RSM) was used to determine the

optimum gasification condition. Moreover, an economic analysis model was presented to

estimate the hydrogen production cost from the gasification of CS and PKS. The knowledge

obtained through this study is extremely valuable for exploring thermochemical reactions

between two potential types of biomass and their gasification performances.

2. Materials and methods

2.1. Raw materials and characterization

The CS and the PKS were respectively obtained from a local grocery store and a palm oil

factory in Teluk Intan, Perak, Malaysia. Both shells were then crushed using a professional lab

jaw crusher (Blue Beauty) and sieved using a stainless steel mesh with different mesh sizes to

obtain particle sizes within the ranges of 1–3 mm, 4–7 mm, and 8–11 mm, as presented in Fig.

1. These particle size ranges were chosen because they nearly equal the industrial

recommended ranges for downdraft gasifiers (≤ 10 mm) [24]. The feedstocks were dried in an

oven at 105 ºC for 24 h to ensure the removal of moisture content (< 15%). The proximate and

ultimate analyses were carried out to develop the empirical formula and to examine the

quantitative energy content of the feedstocks. The proximate analysis was performed by

calculating the weight loss of the heated sample in an oven, according to the American Society

for Testing Materials (ASTM) D3172–D3175 method. The ultimate analysis was determined

using a Leco TruSpec CHNS analyzer according to ASTM D3176-09 standard procedure. An

AC-350 bomb calorimeter was used to measure the gross calorific value of the feedstocks,

which is also known as the HHV according to ASTM D1989 standard.

5
ACCEPTED MANUSCRIPT

Fig. 1. Different particle sizes of (a) CS and (b) PKS with the average particle sizes of

8–11 mm (large), 4–7 mm (medium), and 1–3 mm (small).

2.2. Experimental setup

A batch type downdraft fixed-bed with an electrically heated reactor was made from a

stainless steel tube with a diameter of 5.08 cm and a height of 50 cm. A ceramic electrical type

was used to supply the desired heat for start-up and maintain heat loss during the operation.

The operating temperature was measured by an external PID controller coupled with a K-type

thermocouple mounted on the gasifier reactor. The system was heated up to the desired

operation temperature of 700 °C at start-up. The system operating temperature was then

increased to 800 °C and 900 °C. Air, as a gasification agent, was generated from a 20-gallon

air compressor, and the compressed air was introduced into the reactor. Rather than steam or a

mixture of air and steam, air was chosen as the gasifying agent to avoid additional costs for

steam generation. An equivalent ratio (ER) of 2.2 was applied for CS and PKS gasification

because it is the optimum condition for the downdraft reactor. As the desired temperature was

reached, a certain amount of feedstock was fed through the open top of the gasifier. The

6
ACCEPTED MANUSCRIPT

reproducibility of the gasification results was evaluated by repeating three sets of experiments

under similar operational conditions.

Fig. 2. Schematic representation of the batch-type downdraft fixed-bed reactor.

2.3. Sampling and analysis of gasification products

The raw gases were cooled down in six impinger bottles containing isopropanol (99.9%)

for separating condensable components to trap liquid (tar and water) and acquire clean syngas.

The impinger bottles were placed in water and ice–salt mixture at temperatures of 20 °C and -

20 °C, respectively, as shown in Fig. 2. A gas meter and a pump were used to measure and

control the volumetric flow. The clean gas was streamed directly to the Emersion X-Stream

X2GP online gas analyzer with an accuracy of ±0.01% to measure the composition of CO,

CO2, CH4, and H2 gases. The chosen gas analyzer enabled the combination of four channels,

namely, ultraviolet analysis, nondispersive infrared analysis, electrochemical and

paramagnetic oxygen analyses, and thermal conductivity analysis. Moreover, extra syngas was

ignited at the flare point located between the pump and gas cleaning and cooling system.

7
ACCEPTED MANUSCRIPT

After completing each experiment, the isopropanol–tar mixture in the impinger bottles were

mixed, and the tubing was washed with isopropanol. Tar was recovered by evaporating

isopropanol at 45 ℃ in a rotary evaporator under a vacuum of 100 mbar. The tar sampling was

based on the protocol described in CEN/TS 15439 Biomass Gasification-Tar and Particles in

Product Gases-Sampling and Analysis [25]. The solid residue left inside the gasifier was

collected and weighed.

2.4. Calculations of gasification effectiveness

The effectiveness of gasification was evaluated in terms of syngas higher heating value

(HHVsyngas), dry gas yield (Y), carbon conversion (ηc) and cold gas efficiency (η). The syngas

higher heating value (HHVsyngas) was determined by computing the volumetric percentage of

the dry syngas at normal temperature and pressure according to the following equation [26]:

HHVsyngas = (CO vol% × 30.18 + H2 vol% × 30.52 + CH4vol% × 95) × 4.1868 (MJ/Nm3), (1)

where CO vol%, H2 vol%, and CH4 vol% are the average volumetric percentages of CO, CO2,

and CH4 in the syngas, respectively.

The Y was determined using the following equation:

Qa × 0.79
Y = × 100, (2)
Wb (1 - Xash) × N2vol%

where Qa is the flow rate of air (Nm3/h), Wb is the mass flow rate of the biomass (kg/h), Xash is

the ash content in the biomass, and N2 vol% is the nitrogen content in the volumetric percentage

of the dry syngas [26].

The c was obtained according to the following equation:

8
ACCEPTED MANUSCRIPT

Y × (CO vol% + CO2 vol% + CH4 vol%) × 12


ηc = × 100,
22.4 × C %

(3)

where CO vol%, CO2 vol%, and CH4 vol% are the average volumetric percentages of CO, CO2,

and CH4 in the syngas, respectively and C % is the carbon content resulting from the ultimate

analysis of the biomass [26].

The  or gasification system efficiency is obtained according to the following equation:

Hg × Y
η = × %, (4)
Hb

where Hg (MJ/Nm3) and Hb (MJ/kg) represent the HHVsyngas and biomass, respectively. This

equation indicates the chemical energy of the producer gas compared to that of fuel.

2.5. Design of experiments for optimizing gasification performance

RSM is a statistical tool based on the fit of a polynomial equation to the experimental data.

This tool is used to examine the behavior between independent and dependent variables to

obtain optimal response [27]. With the integrated variance (IV)-optimal design as basis, five

replicates at the center point, resulting in a total of 20 experiments, were conducted to optimize

the low and high chosen levels of the independent gasification variables as described in Table

1. In this evaluation, the small (1 mm), medium (6 mm), and large (11 mm) particle sizes were

selected to represent the entire applied particle size range (1–11 mm) for CS and PKS. The IV-

optimal design was selected because it has a satisfactory variance of prediction over the region

of experimentation and tends to generate appropriate criterion for creating 3D graphical

response surfaces [28, 29]. The graphical response surfaces were fitted with a quadratic

polynomial model obtained from Design-Expert® (Stat Ease Inc., USA). This model was the

best fit for the analysis. Temperature and particle size were independent variables, whereas gas

9
ACCEPTED MANUSCRIPT

composition (CO, CO2, CH4, and H2) and gasification performance (HHVsyngas, Y, ηc, and

 were response or dependent variables. The model is significant and near to the experimental

values if P ≤ 0.05.

Table 1. Ranges and levels of different process variables for the IV-optimal design.

Independent variable Range and levels


Low (-1) Center point (0) High (+1)
A - Temperature (°C) 700 800 900
B - Particle Size (mm) 1 6 11

2.6. Mass balance

Mass balance was used to determine the amount of char, which is gas and tar produced

from the gasification of the biomass (CS and PKS). This part is vital to evaluate the overall

thermochemical conversion. In the gasification of the CS, the input flow rates of the CS and

air were 10 g/min and 11 l/min, respectively, whereas in the gasification of the PKS, the input

flow rates of the PKS and air were 10 g/min and 12.6 l/min, respectively. The input air

volumetric flow rates (Qair) for CS and PKS gasification were calculated based on the fixed ER

of 2.2. The input air flow rates differ for CS and PKS gasification because the stoichiometric

ratio of air to fuel (air/fuel) is different, as shown in Table 2. Eqs. 5 and 6 show the total material

balance for gasification input and output, respectively. Eq. 7 was used to measure the mass

flow rate of air [30].

Σ ṁinput = [ṁbiomass + ṁair] input (5)

Σ ṁoutput = [ṁchar + ṁgas + ṁtar] output (6)

ṁair = [Qair × ρair] (7)

where

ṁbiomass is the inlet mass flow rate of the biomass into the gasifier (g/min)

10
ACCEPTED MANUSCRIPT

ṁair is the inlet mass flow rate of air into the gasifier (g/min)

ṁchar is the outlet mass flow rate of the produced char (g/min)

ṁgas is the outlet mass flow rate of the produced gas (g/min)

ṁtar is the outlet mass flow rate of the produced tar (g/min)

Qair is the volumetric flow rate of air (l/min) ρair is the density of air (1.2 kg/m3).

The product gas yield was calculated using Eq. (8), and the mass flow rate for each

produced gas generated during gasification was calculated using Eq. (9) [30].

Wj = V × Z j

(8)

Wj × Mwj
ṁj = 22.4 × ṁbiomass

(9)

where

Wj is the gas yield of each gas (Nm3/kg)

ṁj is the mass flow rate for each produced gas (g/min)

V is the total yield of the produced gas (Nm3/kg)

Zj is the mole fraction of each gas

Mwj is the molar weight of each gas (g/mol)

The total mass flow rate of the produced gas generated (ṁgas) is shown in Eq. (10) [30].

ṁgas = Ʃṁj (10)

11
ACCEPTED MANUSCRIPT

The carbon content in the biomass input was used to determine the inlet mass flow rate of

carbon into the gasifier, as shown in Eq. (11). The total carbon output mass balance resulting

from the conversion of carbon inherent in the biomass input was evaluated using Eq. (12). Air

was assumed as a carbon-free source in the inlet stream. Therefore, only CS and PKS were

considered as carbon source, as shown in Eq. (12) [30].

Σ Ċinput = Čbiomass × ṁbiomass (11)

Σ Ċouput = [Ċchar + Ċgas + Ċtar]output

(12)

where

[Ċbiomass]input is the inlet mass flow rate of carbon (from biomass) into the gasifier (g/min)

Ċchar is the mass flow rate of the produced char (g/min)

Ċgas is the mass flow rate of the produced gas (g/min)

Ċtar is the mass flow rate of the produced tar (g/min)

Čbiomass is the carbon content in the biomass (g)

ṁbiomass is the inlet mass flow rate of the biomass into the gasifier (g/min).

The mass flow rate of carbon from the generated individual carbonaceous species was

calculated using the following equation [30]:

12
ṁj = wj × 22.4 × ṁbiomass (13)

Hence, the total mass flow rate of carbon (Ċgas) from the generated individual carbonaceous

gas species (Ċi) was obtained using the following equation:

12
ACCEPTED MANUSCRIPT

Ċgas = ƩĊi (14)

The mass flow rates of char and tar were obtained using Eqs. 15 and 16.

Ċchar = Čchar × ṁchar/t

(15)

Ċtar = Čtar × ṁtar/t (16)

where

Čchar is the weight percentage of carbon in the produced char

Čtar is the weight percentage of carbon in the produced tar

ṁchar is the mass flow rate of the produced char (g/min)

ṁtar is the mass flow rate of produced tar (g/min)

t is the gasification time (min).

The char and tar carbon analyses were performed using a Leco TruSpec CHNS analyzer

according to ASTM D3176-09 standard procedure. The total error (te) in the material and

carbon mass balances between the input and output streams was evaluated using Eqs. 17 and

18 [30].

∑𝑚𝑖𝑛𝑝𝑢𝑡 - ∑𝑚𝑜𝑢𝑡𝑝𝑢𝑡
Total error (te) = ∑𝑚𝑖𝑛𝑝𝑢𝑡 × 100 % (17)

∑𝐶𝑖𝑛𝑝𝑢𝑡 - ∑𝐶𝑜𝑢𝑡𝑝𝑢𝑡
Total error (te) = ∑𝐶𝑖𝑛𝑝𝑢𝑡 × 100 % (18)

1. Results and discussion

1.1. Analysis of raw materials

13
ACCEPTED MANUSCRIPT

The values of the proximate and ultimate analyses and the HHV of the feedstocks can

decide the characteristics of each feedstock. These values also vary widely depending on the

plant species. Table 2 provides the analysis results of the CS and PKS. According to the

proximate analysis, the CS has a higher amount of moisture (M) (ca. 10.5 wt%) compared to

the PKS (ca. 9.4 wt%). These values are acceptable for thermochemical conversion in a fixed-

bed downdraft reactor. In addition, the CS contains a higher amount of volatile matter (VM)

(ca. 71.1 wt%) compared to the PKS (ca. 69.8 wt%). Biomass with a high-value VM content

helps in simplifying biomass combustion [31]. Table 2 clearly shows that the fixed carbon

content in the CS (17.6 wt%) is higher than that of the PKS (16.8 wt%). This content can be

correlated to the amount of solid carbon left after the biomass undergoes full devolatilization

[32]. Moreover, the CS has a lower ash (A) content (0.80 wt%) than the PKS (4.00 wt%),

indicating that the CS has a higher conversion quality with lower slag [33, 34].

The carbon (C) and hydrogen (H) contents in the CS are slightly higher than those in the

PKS in the ultimate analysis. The high C and H contents indicate an HHV signature due to a

higher energy level inherent in carbon–carbon bonds compared to that in carbon–hydrogen and

carbon–oxygen bonds [35]. The O content in the PKS is slightly higher than that in the CS. A

high O content may reduce the energy content of the biomass because it is an unsupported

element for combustion [36]. In addition, the CS and PKS show the presence of nitrogen (N)

and sulphur (S) contents in the produced gas. These gases are harmful to the environment, so

additional treatment is needed during gasification. The ultimate analysis reveals that the C, H,

O, N, and S contents in the PKS and CS are in agreement with other reported values [37–40].

According to the analysis results, the CS has better characteristics for producing higher quality

syngas than the PKS.

Table 2. Characterization of the coconut shell (CS) and palm kernel shell (PKS).

14
ACCEPTED MANUSCRIPT

Biomass
Characteristics Parameter
CS PKS
Moisture, M 10.5 9.40
Volatile matter, VM 71.1 69.8
Proximate analysisa (wt%)
Fixed carbon, FC 17.6 16.8
Ash, A 0.80 4.00
Carbon, C 48.6 46.3
Ultimate analysisb Hydrogen, H 5.97 5.72
(wt%) Oxygenc, O 43.8 47.6
Nitrogen, N 0.62 0.70
Sulfur, S 1.09 0.64
Higher heating value, HHV (MJ/kg) 18.8 17.2
Empirical formula CH1.465O0.677N0.0123 CH1.5478O0.711N0.0121
Stoichiometric air/fuel ratio (kg/kg) 5.006 5.754
a As received
b Dry basis
c O wt% = 100 wt% - C wt% - H wt% - N wt% - S wt%

1.2. Effect of particle size and temperature on gas composition

The produced gases during gasification mainly consist H2, CO, CO2, H2O, CH4, C2+, and

N2 [41]. The produced gas composition varies widely according to the employed operating

conditions during gasification. Such variations in the gas composition can be easily expressed

by reactions shown in the following equations [17, 42, 43].

CH4 + H2O ↔ CO + 3H2 ∆H = +206kJ/mol (19)

CH4 + H2O ↔ CO2 + 4H2 ∆H = +165kJ/mol (20)

CH4 + CO2 ↔ 2CO + 2H2 ∆H = +247kJ/mol (21)

CnHm(tar) + nH2O ↔ (m/2)H2 + nCO ∆H > 0 (22)

CnHm(tar) + nC2O ↔ (m/2)H2 + 2nCO ∆H > 0 (23)

CnHm(tar) ↔ (m/2)H2 + nC ∆H > 0 (24)

C + H2O ↔ CO + H2 ∆H = 131kJ/mol (25)

C + CO2 ↔ 2CO ∆H = 172kJ/mol (26)

CO + H2O ↔ CO2 + H2 ∆H =41:98kJ/mol (27)

15
ACCEPTED MANUSCRIPT

Figs. 3 and 4 depict the 3D surface responses of H2, CO, CO2, and CH4 vol% generated

from the integration effect of particle size and temperature during CS and PKS gasification

with an extremely low P-value (P < 0.0001). High concentrations of H2, CO, and CH4 can be

observed in the small particle size, and these trends may be attributed to the tar cracking and

reforming reactions (Eqs. 22– 24), the water–gas shift reaction (Eq. 25), the Boudouard

reaction (Eq. 26), and the exothermic water–gas shift reaction (Eq. 27) [21]. A similar trend

was also obtained by other studies using particle sizes ranging from 0.3–1 mm, 5–50 mm, and

0.075–1.2 mm in a fluidized-bed reactor, a downdraft gasifier, and an updraft fixed-bed reactor,

respectively [16, 23, 44]. Patel et al. [22] studied the effect of various lignite particle sizes on

the performance of a downdraft gasifier. The particle sizes in the study varied from small (13–

16 mm) to large (28–31 mm). The authors noticed that gasification using particle sizes ranging

from 22–25 mm showed the highest H2 and CO compositions.

The CO2 produced from the CS and the PKS showed a trend in the opposite direction,

wherein it slightly increased as the particle size increased, as depicted in Figs. 4a and 4b. This

result was in agreement with other studies [16, 17, 23]. However, a different trend in CO2

composition was observed by Yin et al. [20] during gasification of peach pruning with five

different particle size fractions (< 1, 1–2, 2–4, 4–6, and 6–8 cm) in a pilot scale downdraft

fixed-bed reactor. The authors claimed that the CO2 percentage decreased from 14.36% to

8.28% when the particle size was increased from < 1 cm to 4–6 cm, respectively. The obtained

results also indicated a slightly higher H2 and CH4 composition for the CS compared to those

for the PKS. This result may be due to the higher moisture and carbon content of the CS

compared to the PKS, as shown in Table 1. A few studies claimed that the high H2 value

resulted from the high moisture content value [11, 43]. The high O and low moisture content

in the PKS tended to generate more CO and CO2 rather than H2.

16
ACCEPTED MANUSCRIPT

Temperature plays a critical role in the overall biomass gasification process. As shown in

Figs. 3a, 3b, 3c, and 3d, the H2 and CO percentages for the CS and the PKS increased drastically

when the temperature increased from 700 °C to 900 °C. The gas species produced from

gasification of the CS and the PKS may undergo several reactions at a high reactor temperature,

such as endothermic methane steam reforming (Eqs. 19 and 20) and dry reforming reactions

(Eq. 21) [45]. Furthermore, secondary reactions, such as tar reforming and cracking also

contributed to an increase in H2 percentage and other incondensable gases (Eqs. 22– 24) [16].

The same incremental trend in the H2 percentage with increasing temperature was observed in

the research conducted by Ref. [46, 47]. The remarkable increase in CO composition with

increasing temperature may be due to the conversion of CO2 to CO via methane dry reforming

(Eq. 21), tar cracking (Eq. 23), and the Boudouard reaction (Eq. 26), as discussed by Hernández

et al. [48].

The effect of temperature on the CO2 composition of the CS and PKS clearly showed a

completely different trend compared to the trends of other gas compositions. As shown in Figs.

4a and 4b, CO2 composition increased from 700 °C to 800 °C before drastically dropping at

temperatures above 800 °C. The phenomenon of high CO2 composition at low temperatures (<

800 °C) was due to the influence of the water–gas shift reaction (Eq. 24), whereas at high

temperatures (> 800 °C), the Boudouard reaction becomes dominant (Eq. 23), leading to a

reduction in the CO2 composition [11, 49].

The CH4 composition for the CS and PKS in Figs. 4c and 4d shows a parabolic decreasing

trend as temperature increases. This finding is consistent with other reported studies on bamboo

and PKSs [19, 50]. At 900 °C with the biggest particle size, the CS and PKS reveal the lowest

CH4 composition, which is 2.81% and 2.39%, respectively. Based on the obtained results, the

effect of temperature on gas composition showed the same phenomena as the effect of particle

size, wherein the CS H2 and CH4 percentages were slightly higher whereas the presented CO

17
ACCEPTED MANUSCRIPT

and CO2 percentages were slightly lower than those of the PKS. As explained earlier, the

moisture, C, and O contents in raw materials play an important role in remarkably influencing

the composition of valuable gases.

Fig. 3. 3D surface response of the gas composition for the CS: (a) H2 and (c) CO, and for the

PKS: (b) H2 and (d) CO.

18
ACCEPTED MANUSCRIPT

Fig. 4. 3D surface response of the gas composition for the CS: (a) CO2 and (c) CH4, and for

the PKS: (b) CO2 and (d) CH4.

1.3. Effect of particle size and temperature on gasification performance

Figs. 5 and 6 illustrate the 3D surface responses of HHVsyngas, Y, c, and  from the

integrated effects of particle size and temperature during CS and PKS gasification with a

significantly low P-value (P < 0.0001). The HHVsyngas, Y, c, and  for the CS and PKS

gradually increases as the particle size decreases and temperature increases, respectively. This

variation is because smaller particle sizes tend to have a higher rate of thermochemical reaction

than bigger particle sizes which enhance the H2, CO, and CH4 percentages and thus leads to

better HHVsyngas. However, Li et al. [19] reported an increase in HHVsyngas content from 8.99

MJ/Nm3 to 10.28 MJ/Nm3 during the gasification of palm oil wastes (shell, fiber, and empty

fruit bunches [EFB]) when the particle size was increased from < 0.15 mm to 5 mm.

A slightly high Y trend is observed from gasification of small CS and PKS particle sizes.

A similar trend is also reported for gasification of EFBs and pine sawdust with particle sizes

ranging from 0.5–1.0 mm and 0.075–1.2 mm using fluidized-bed and fixed-updraft reactors,

respectively [16, 18]. Small particle sizes have large surface areas and tend to generate high Y

because of their fast combustion rate [43]. However, the Y of the CS is slightly higher than that

of the PKS because the generated gas product for the CS is slightly higher than that of the PKS.

The c and  for the CS and PKS decrease as particle size increases. The decrease in c is

due to the low carbonaceous gas values generated during gasification. However, the c for the

PKS is higher than that of the CS as a result of high CO, CO2, and CH4 concentrations in the

produced gas. The drop in the  is due to the integration effect from the low HHVsyngas and Y

19
ACCEPTED MANUSCRIPT

values at big particle sizes, as the  is the ratio of the syngas heating value (Hg) to the heating

value of biomass (Hb), as shown in Eq. 4. However, a slightly higher  is obtained from the CS

compared to that of the PKS with the smallest particle size. This result can be explained by the

high HHVsyngas and Y values of the produced gas at the smallest CS particle size. Inayat et al.

[47] tested the gasification performance of an oil palm frond and CS blending in ratios of 100:0,

80:20, and 60:40 with particle sizes ranging from 10–25 mm in a downdraft gasifier. According

to the study, the Y, c, and  are notably higher than those obtained from gasification of the CS

in the present study. This difference is expected because of the different gasification conditions,

particle sizes, and biomass types used in the studies.

As seen in Figs. 5 and 6, the operating temperature shows a greater effect on the gasification

performance of the CS and the PKS compared to particle size. As mentioned earlier, the H2 and

CO percentage concentrations are mainly dependent on the operating temperature of the

gasifier, which notably reflects the HHVsyngas value. Ariffin et al. [51] also observed the same

HHVsyngas phenomena when increasing the temperature of a medium-scale downdraft gasifier

from 399 °C to 700 °C during PKS gasification. However, the reported values of the HHVsyngas

(4.45–4.89 MJ/Nm3) are slightly higher than those of the PKS reported in this study. This

difference can be correlated to the variation in gasification conditions using the downdraft

gasifier in this current study. The result also indicates that the Y and the  for the CS are slightly

higher than those of the PKS due to the higher flow rate of the generated gas and the integration

effects of the Y, Hg, and Hb, as shown in Eq. 4.

The influence of gasification temperature on the HHVsyngas, Y, c, and  is also in

agreement with the findings of numerous researchers, wherein these parameters remarkably

increase with temperature [52–54]. Lahijani et al [45] investigated the effect of temperature

(650–1050 °C) on the gasification performance of EFB and sawdust in a bubbling fluidized

bed reactor. The authors found that when the temperature was increased, the HHVsyngas and the

20
ACCEPTED MANUSCRIPT

Y increased from 3.27 MJ/Nm3 to 5.37 MJ/Nm3 and from 1.36 Nm3/kg to 2.10 Nm3/kg for the

EFB and from 3.86 MJ/Nm3 to 5.87 MJ/Nm3 and from 1.28 MJ/Nm3 to 1.95 Nm3/kg for the

sawdust, respectively.

Fig. 5. 3D surface response of gasification performance for the CS: (a) HHVsyngas and (c) Y

and for the PKS: (b) HHVsyngas and (d) Y.

21
ACCEPTED MANUSCRIPT

Fig. 6. 3D surface response of gasification performance for the CS: (a) c and (c)  and for

the PKS: (b) c and (d) 

1.4. Effect of particle size and temperature on mass balance

The effect of particle size and temperature on material and carbon mass balances from

gasification of the CS and PKS are evaluated and presented in Tables 3 and 4, respectively.

The mass flow rate of the output gas (ṁgas) decreases when the particle size of the CS and PKS

is increased. By contrast, the mass flow rates of char (ṁchar) and tar (ṁtar) increase when the

particle size of the CS and PKS is increased. Based on the carbon balance, the total carbon

output increases as particle size decreases. This behavior is due to the reduction of temperature

gradient inside small particles with high surface areas, resulting in better reaction rates during

gasification than that of particles with a bigger size. The total material and carbon mass balance

errors increase with increasing particle size because of the decrease in the flow rate of the

output gas. A similar trend is also shown when the operating temperature is increased from 700

°C to 900 °C for all particle sizes. However, the influence of temperature on the gasification

22
ACCEPTED MANUSCRIPT

reaction is more significant than that of particle size, as confirmed from the reduced total

material and carbon mass balance errors at high operating temperatures for each particle size

of the biomass. Carpenter et. al [55] also found a decrease in the total carbon mass balance

error during gasification of corn stover, Vermont mixed wood, switch grass, and wheat straw

when the temperature of a hot fluidized bed gasifier was increased from 600 °C to 850 °C.

Abdoulmoumine et. al [56] observed a similar decreasing trend of material mass balance error

from gasification of loblolly pine when the temperature of a fluidized bed gasifier was

increased from 790 °C to 1078 °C. The thermal decomposition of heavy hydrocarbons, such as

tar and char, are enhanced at high operating temperatures, which lead to a high flow rate of the

output gas. Moreover, high operating temperatures improve the conversion of char and tar into

light gases through the Boudouard and tar-cracking reactions [44, 46]. This phenomenon

explains why the total material mass balance error is smaller at higher operating temperatures

compared to that at lower operating temperatures. The reduced total carbon mass error at high

operating temperatures further confirms the improved conversion of char and tar into light

gases. This conversion results in a low amount of carbon in the char and tar produced at high

temperatures.

Gasification of the CS and PKS at 900 °C with particle sizes ranging from 1–3 mm shows

the highest output gas flow rate and the lowest char and tar flow rates. However, gasification

of the CS shows a slightly higher output gas with a lower material mass balance error than that

of the PKS. This result is because the high moisture and carbon contents in the CS tend to

produce additional valuable gases, as presented in Table 1. This result is also consistent with

the 3D surface response plots shown in Fig. 3. From the overall material and carbon mass

balance, the total material and carbon mass errors detected during gasification range from

13.03–16.84% and 5.01–10.49% for the CS and 15.55–20.60% and 9.41–14.18% for the PKS,

respectively. These errors are considerably high because the water produced from the

23
ACCEPTED MANUSCRIPT

gasification reaction was neglected in the material mass balance evaluation. The errors can also

be correlated to the limitation of the gas analyzer as it is not designed to detect gases that are

heavier than CH4, such as C2H2, C2H4, and C2H6.

24
Table 3. Overall material mass balance for gasification of the CS and the PKS.

Particle size Coconut shell (CS) Palm kernel shell (PKS)


Material source
range (mm) Temperature (°C) Temperature (°C)
700 800 900 700 800 900
Fixed mass flow rate of input air and biomass 1-3 23.2 23.2 23.2 25.1 25.1 25.1
(Σ ṁinput) 4-7 23.2 23.2 23.2 25.1 25.1 25.1
(g/min) 8-11 23.2 23.2 23.2 25.1 25.1 25.1

Mass flow rate of output gas 1-3 15.57 ± 0.09 16.69 ± 0.10 17.22 ± 0.04 15.98 ± 0.15 17.03 ± 0.10 17.92 ± 0.09
(ṁgas) 4-7 15.52 ± 0.08 16.57 ± 0.12 17.10 ± 0.10 15.86 ± 0.13 16.98 ± 0.12 17.87 ± 0.10
(g/min) 8-11 15.48 ± 0.09 16.45 ± 0.15 17.02 ± 0.12 15.63 ± 0.15 16.90 ± 0.15 17.81 ± 0.12

Mass flow rate of output char 1-3 3.29 ± 0.06 3.11 ± 0.06 2.86 ± 0.07 3.66 ± 0.06 3.37 ± 0.09 3.16 ± 0.04
(ṁchar) 4-7 330 ± 0.05 3.13 ± 0.04 2.88 ± 0.04 3.71 ± 0.05 3.40 ± 0.08 3.18 ± 0.09
(g/min) 8-11 3.32 ± 0.07 3.16 ± 0.05 2.90 ± 0.10 3.80 ± 0.06 3.45 ± 0.17 3.20 ± 0.07

Mass flow rate of output tar 1-3 0.46 ± 0.03 0.19 ± 0.01 0.10 ± 0.01 0.49 ± 0.01 0.26 ± 0.02 0.11 ± 0.01
(ṁtar) 4-7 0.47 ± 0.03 0.21 ± 0.01 0.12 ± 0.02 0.50 ± 0.01 0.27 ± 0.02 0.12 ± 0.01
(g/min) 8-11 0.48 ± 0.02 0.23 ± 0.01 0.13 ± 0.01 0.51 ± 0.01 0.28 ± 0.01 0.13 ± 0.01

Total mass flow rate of output 1-3 19.4 ± 0.20 20.0 ± 0.10 20.2 ± 0.20 20.1 ± 0.19 20.7 ± 0.28 21.3 ± 0.20
(Σ ṁoutput) 4-7 19.3 ± 0.08 19.9 ± 0.15 20.1 ± 0.12 20.0 ± 0.12 20.6 ± 0.19 21.2 ± 0.10
(g/min) 8-11 19.2 ± 0.09 19.8 ± 0.15 20.0 ± 0.14 19.9 ± 0.10 20.5 ± 0.08 21.1 ± 0.14

Total material mass error 1-3 16.65 ± 0.17 13.83 ± 0.81 13.03 ± 0.41 19.74 ± 0.76 17.62 ± 1.10 15.55 ± 0.40
((Σ ṁinput – Σ ṁoutput) / Σ ṁinput) × 100 4-7 16.78 ± 0.11 14.19 ± 0.62 13.42 ± 0.55 20.07 ± 0.49 17.71 ± 0.74 15.63 ± 0.67
(%) 8-11 16.84 ± 0.10 14.49 ± 0.69 13.56 ± 0.60 20.60 ± 0.22 17.80 ± 03.0 15.71 ± 0.16

25
Table 4. Overall carbon mass balance for gasification of the CS and the PKS.

Coconut shell (CS) Palm kernel shell (PKS)


Material source Particle size (mm)
Temperature (°C) Temperature (°C)
700 800 900 700 800 900
Mass flow rate of input biomass carbon 1-3 4.86 4.86 4.86 4.63 4.63 4.63
(Σ Ċinput) 4-7 4.86 4.86 4.86 4.63 4.63 4.63
(g/min) 8-11 4.86 4.86 4.86 4.63 4.63 4.63

Mass flow rate of output gas carbon 1-3 1.80 ± 0.04 1.90 ± 0.05 1.93 ± 0.02 1.79 ± 0.03 1.90 ± 0.02 1.92 ± 0.03
(Ċgas) 4-7 1.78 ± 0.02 1.89 ± 0.02 1.92 ± 0.01 1.76 ± 0.03 1.89 ± 0.01 1.91 ± 0.02
(g/min) 8-11 1.77 ± 0.01 1.87 ± 0.01 1.90 ± 0.01 1.75 ± 0.02 1.87 ± 0.01 1.90 ± 0.01

Mass flow rate of output char carbon 1-3 2.34 ± 0.08 2.36 ± 0.09 2.41 ± 0.07 2.34 ± 0.09 2.37 ± 0.10 2.42 ± 0.10
(Ċchar) 4-7 2.33 ± 0.06 2.35 ± 0.07 2.40 ± 0.06 2.33 ± 0.06 2.36 ± 0.09 2.41 ± 0.07
(g/min) 8-11 2.31 ± 0.05 2.34 ± 0.05 2.39 ± 0.05 2.31 ± 0.05 2.35 ± 0.06 2.40 ± 0.06

Mass flow rate of output tar carbon 1-3 0.11 ± 0.03 0.07 ± 0.03 0.05 ± 0.04 0.12 ± 0.04 0.09 ± 0.03 0.05 ± 0.05
(Ċtar) 4-7 0.10 ± 0.02 0.06 ± 0.02 0.04 ± 0.03 0.11 ± 0.03 0.08 ± 0.02 0.05 ± 0.04
(g/min) 8-11 0.09 ± 0.02 0.05 ± 0.01 0.04 ± 0.01 0.10 ± 0.02 0.07 ± 0.01 0.05 ± 0.03

Total mass flow rate of output carbon 1-3 4.24 ± 0.09 4.33 ± 0.01 4.40 ± 0.05 4.25 ± 0.04 4.35 ± 0.05 4.40 ± 0.26
(Σ Ċoutput) 4-7 4.20 ± 0.06 4.31 ± 0.03 4.37 ± 0.01 4.21 ± 0.06 4.33 ± 0.48 4.37 ± 0.21
(g/min) 8-11 4.17 ± 0.02 4.28 ± 0.04 4.34 ± 0.03 4.17 ± 0.03 4.30 ± 0.57 4.36 ± 0.56

Total carbon mass error 1-3 8.24 ± 1.96 6.92 ± 0.25 5.01 ± 1.04 12.23 ± 0.09 10.81 ± 1.09 9.41 ± 0.57
((Σ Ċinput – Σ Ċoutput) / Σ Ċinput) × 100 4-7 9.61 ± 1.34 7.38 ± 0.66 6.02 ± 0.24 13.57 ± 1.38 11.32 ± 1.03 10.09 ± 0.46
(%) 8-11 10.49 ± 0.48 7.85 ± 0.89 6.58 ± 0.77 14.18 ± 0.58 11.98 ± 1.24 10.74 ± 1.20

26
ACCEPTED MANUSCRIPT

1.5. Economic analysis of CS and PKS gasification

The downdraft reactor used in this study is designed only for low biomass flow rates (0.6

kg/h) for H2 production. Therefore, the scaling-up of the existing reactor for high biomass flow

rates is necessary to estimate a reliable H2 production cost. The details of the optimum operating

conditions (particle size of 1 mm and operating temperature of 900 °C) for the scaled-up

gasifier (6 kg/h) with a high biomass flow rate is estimated and presented in Table 5. The H2

production rate per day (rḣ₂) was estimated using Eq. 28.

rh2 = ḃ × Ù × ḣn (28)

where

ḃ is the biomass feed rate (kg/h)

Ù is the H2 production rate (kg/d)

ḣn is the hydrogen generation rate (kg H2/kg biomass).

The assumption of capital cost for this economic analysis consists of the equipment, operating,

and maintenance costs and interest, as presented in Table 6. According to the scale-up

specifications, the downdraft gasifier can generate 3.528 kg H2/d and 3.2832 kg H2/d, using the

CS and PKS, respectively.

Table 7 compares the H2 production cost (US$/Kg) with other production processes. The

cost for H2 production per Kg biomass (Cb) was calculated using Eq. 29.

P
Cb = Ha (29)

where

P is the capital cost (RM/yr) and Ha is the H2 production per year (kg/yr).

27
ACCEPTED MANUSCRIPT

The estimated H2 production cost from this analysis is US$3.00/kg and US$3.23/kg for the CS

and the PKS, respectively. The result indicates that the H2 production cost generated from this

study is comparable with other gasification processes, as shown in Table 6. However, the H2

production cost in this study is slightly higher compared to that reported by Muhammed et al.

[16] using the same biomass flow rate amount. This difference is closely related to the type of

gasifier and materials used during the gasification. The current reactor design has the potential

to be used as an alternative process for low cost H2 production.

Table 5. Operating conditions for gasification of the CS and PKS using a downdraft reactor.

Material Coconut shell (CS) Palm kernel shell (PKS)


Biomass Feed Rate (kg/h), ḃ 6 6
Hydrogen generation rate (Kg H2/kg
0.0245 0.0228
biomass), ḣn
H2 production rate (kg/d), Ù 144 144
Air flow (Nm3/h) 7.92 9.07
Temperature in the gasifier (°C) 900 900
Particle size (mm) 1 1
H2 (vol%)* 14.6 13.3
CO (vol%)* 17.4 17.7
CO2 (vol%)* 14.8 15.7
CH4 (vol%)* 3.42 2.75
Syngas HHV (MJ/m3)* 5.39 5.08
* The optimum results obtained from RSM

Table 6. Cost analysis for the gasification of the CS and PKS using a downdraft reactor.

System and typical investment Detail Unit CS cost PKS cost Reference
Biomass Current
Operating system capacity Kg/yr 52560 52560
consumption Study
H2 production per Current
Kg/yr 1287.72 1198.37
year, Ha Study
Fixed-bed
Initial cost downdraft gasifier RM 8984 8984 [16]
system
Construction RM 1746 1746 [16]
Feedstock
Operating cost RM/yr 2592 2592 [16]
(RM 50/tonne)
Electricity
RM/yr 1095 1095
(RM 0.23/kWh)
10% of initial
Interest cost RM/yr 1073 1073 [16]
cost/yr
Maintenance cost 5% of initial cost/yr RM/yr 536.5 536.5 [16]
Total annual cost Capital cost, p RM/yr 16026.5 16026.5
Current
Cost of H2, Cb RM/Kg 12.44 13.77
study
US$/Kg 3.00 3.23

28
ACCEPTED MANUSCRIPT

Table 7. Comparison of hydrogen production costs through different processes.

Biomass
H2 Cost
Feed rate Process applied Material Gasifier type Reference
(US$/Kg)
(kg/h)
Biomass air-steam gasification Japanese
4.60 450 Moving bed [60]
with multiple heat cedar
Biomass pyrolysis for hydrogen
production using CO-shift Pyrolysis
4.28 4166 Wood [61]
combined with high pressure Furnace
steam
Empty Fluidized
2.11 6.0 Biomass air gasification [62]
fruit bunch bed
Biomass steam gasification with Oil palm Fluidized
1.91 0.8 [63]
in situ CO2 capture using CaO waste bed
Biomass oxygen-rich air Fixed-bed
1.69 266.7 Rice husk [64]
gasification with CO shift downdraft
Coconut Fixed-bed Current
3.00 6.0 Biomass air gasification
shell downdraft Study
Palm
Fixed-bed Current
3.23 6.0 Biomass air gasification kernel
downdraft Study
shell

4. Conclusions

Air gasification of a CS and a PKS using various particle sizes (1–3, 4–7, and 8–11 mm) at

various operating temperatures (700, 800, and 900 °C) was successfully conducted in a batch-

type downdraft fixed-bed gasifier. The gas composition increased when the particle size and

temperature were reduced and increased, respectively. The CS produced more valuable gases

(H2 and CO) than the PKS under all conditions. Similarly, the gasification performance of the

HHVsyngas, Y, c, and  for the CS was slightly higher than that for the PKS at all conditions.

These values increased when the particle size and temperature decreased and increased,

respectively. The amount of gas produced based on the material mass balance showed an

increasing trend, whereas the amount of char and tar produced showed a decreasing trend when

the particle size and temperature were decreased and increased, respectively. The predicted

cost for H2 production using the CS (US$3.00/kg) was slightly lower than that of the PKS

(US$3.23/kg). Moreover, temperature has a greater role than particle size in influencing the

gasification reaction rate. The integration of high temperature and small particle size may

enhance the gasification performance and lead to high Y with less char and tar contents. Overall,

29
ACCEPTED MANUSCRIPT

the gasification performance using the CS was slightly better than that of the PKS at all

conditions. However, the CS and the PKS demonstrated considerable potential as alternative

sources for future gasification feedstock.

Acknowledgements

The authors gratefully acknowledge the Research University Grant (GUP-2018-040) provided

by the Universiti Kebangsaan Malaysia (UKM). Ahmad Zubair Yahaya thankfully

acknowledges the Ministry of Education Malaysia for the MyPhD scholarship. The facility

supports from the Centre for Research and Instrumentation Management (CRIM) of UKM and

Universiti Teknologi PETRONAS are also gratefully acknowledged.

References

[1] EIA. International Energy Outlook 2017. Int Energy Outlook 2017:151.
[2] Oilseeds and Oils Report 2015 n.d.
[3] Alipour Moghadam Esfahani R, Osmieri L, Specchia S, Yusup S, Tavasoli A,
Zamaniyan A. H2-rich syngas production through mixed residual biomass and HDPE
waste via integrated catalytic gasification and tar cracking plus bio-char upgrading.
Chem Eng J 2017;308:578–87. doi:10.1016/j.cej.2016.09.049.
[4] Nanda S, Isen J, Dalai AK, Kozinski JA. Gasification of fruit wastes and agro-food
residues in supercritical water. Energy Convers Manag 2016;110:296–306.
doi:10.1016/j.enconman.2015.11.060.
[5] Widjaya ER, Chen G, Bowtell L, Hills C. Gasification of non-woody biomass: A
literature review. Renew Sustain Energy Rev 2018;89:184–93.
doi:10.1016/j.rser.2018.03.023.
[6] Samiran NA, Mohd Jaafar MN, Chong CT, Jo-Han N. A review of palm oil biomass as
a feedstock for syngas fuel technology. J Teknol 2015. doi:10.11113/jt.v72.3932.
[7] Valdés CF, Marrugo G, Chejne F, Montoya JI, Gómez CA. Pilot-scale fluidized-bed co-
gasification of palm kernel shell with sub-bituminous coal. Energy and Fuels
2015;29:5894–901. doi:10.1021/acs.energyfuels.5b01342.
[8] Shahbaz M, Yusup S, Inayat A, Patrick DO, Ammar M, Pratama A. Cleaner Production
of Hydrogen and Syngas from Catalytic Steam Palm Kernel Shell Gasification Using

30
ACCEPTED MANUSCRIPT

CaO Sorbent and Coal Bottom Ash as a Catalyst. Energy and Fuels 2017;31:13824–33.
doi:10.1021/acs.energyfuels.7b03237.
[9] Jeya Singh VC, Sekhar SJ. Performance studies on a downdraft biomass gasifier with
blends of coconut shell and rubber seed shell as feedstock. Appl Therm Eng 2016;97:22–
7. doi:10.1016/j.applthermaleng.2015.09.099.
[10] Kumar A, Jones DD, Hanna MA. Thermochemical biomass gasification: A review of
the current status of the technology. Energies 2009;2:556–81. doi:10.3390/en20300556.
[11] Emami Taba L, Irfan MF, Wan Daud WAM, Chakrabarti MH. The effect of temperature
on various parameters in coal, biomass and CO-gasification: A review. Renew Sustain
Energy Rev 2012;16:5584–96. doi:10.1016/j.rser.2012.06.015.
[12] Jangsawang W, Laohalidanond K, Kerdsuwan S. Optimum Equivalence Ratio of
Biomass Gasification Process Based on Thermodynamic Equilibrium Model. Energy
Procedia, vol. 79, 2015, p. 520–7. doi:10.1016/j.egypro.2015.11.528.
[13] Ruiz JA, Juárez MC, Morales MP, Muñoz P, Mendívil MA. Biomass gasification for
electricity generation: Review of current technology barriers. Renew Sustain Energy
Rev 2013;18:174–83. doi:10.1016/j.rser.2012.10.021.
[14] Di Blasi C, Branca C. Temperatures of wood particles in a hot sand bed fluidized by
nitrogen. Energy and Fuels 2003;17:247–54. doi:10.1021/ef020146e.
[15] Mohd Salleh MA, Nsamba HK, Yusuf HM, Idris A, Ghani WAWAK. Effect of
Equivalence Ratio and Particle Size on EFB Char Gasification. Energy Sources, Part A
Recover Util Environ Eff 2015;37:1647–62. doi:10.1080/15567036.2011.555440.
[16] Mohammed MAA, Salmiaton A, Wan Azlina WAKG, Mohammad Amran MS,
Fakhru’L-Razi A. Air gasification of empty fruit bunch for hydrogen-rich gas
production in a fluidized-bed reactor. Energy Convers Manag 2011;52:1555–61.
doi:10.1016/j.enconman.2010.10.023.
[17] Lv PM, Xiong ZH, Chang J, Wu CZ, Chen Y, Zhu JX. An experimental study on
biomass air-steam gasification in a fluidized bed. Bioresour Technol 2004;95:95–101.
doi:10.1016/j.biortech.2004.02.003.
[18] Luo S, Xiao B, Guo X, Hu Z, Liu S, He M. Hydrogen-rich gas from catalytic steam
gasification of??biomass in a fixed bed reactor: Influence of particle size on gasification
performance. Int J Hydrogen Energy 2009;34:1260–4.
doi:10.1016/j.ijhydene.2008.10.088.
[19] Li J, Yin Y, Zhang X, Liu J, Yan R. Hydrogen-rich gas production by steam gasification
of palm oil wastes over supported tri-metallic catalyst. Int J Hydrogen Energy

31
ACCEPTED MANUSCRIPT

2009;34:9108–15. doi:10.1016/j.ijhydene.2009.09.030.
[20] Yin R, Liu R, Wu J, Wu X, Sun C, Wu C. Influence of particle size on performance of
a pilot-scale fixed-bed gasification system. Bioresour Technol 2012;119:15–21.
doi:10.1016/j.biortech.2012.05.085.
[21] Luo S, Xiao B, Hu Z, Liu S, Guan Y, Cai L. Influence of particle size on pyrolysis and
gasification performance of municipal solid waste in a fixed bed reactor. Bioresour
Technol 2010;101:6517–20. doi:10.1016/j.biortech.2010.03.060.
[22] Patel VR, Upadhyay DS, Patel RN. Gasification of lignite in a fixed bed reactor:
Influence of particle size on performance of downdraft gasifier. Energy 2014;78:323–
32. doi:10.1016/j.energy.2014.10.017.
[23] Inayat M, Sulaiman SA, Kumar A, Guangul FM. Effect of fuel particle size and blending
ratio on syngas production and performance of co-gasification. J Mech Eng Sci
2016;10:2188–200. doi:10.15282/jmes.10.2.2016.21.0205.
[24] Wuxi Teneng Power Machinery Co., Ltd. n.d. http://www.powermaxgasifier.com/
(accessed March 8, 2019).
[25] van de Kamp W, de Wild P, Knoef HAM, Neeft JPA, Kiel JHA. Tar measurement in
biomass gasification , standardisation and supporting R & D 2006:1–168.
[26] Xiao R, Zhang M, Jin B, Huang Y, Zhou H. High-temperature air/steam-blown
gasification of coal in a pressurized spout-fluid bed. Energy and Fuels 2006;20:715–20.
doi:10.1021/ef050233h.
[27] Qin X, Wang Y, Lu C, Huang S, Zheng H, Shen C. Structural acoustics analysis and
optimization of an enclosed box-damped structure based on response surface
methodology. Mater Des 2016. doi:10.1016/j.matdes.2016.04.063.
[28] Anderson MJ, Whitcomb PJ. Practical Aspects for Designing Statistically Optimal
Experiments. J Stat Sci Appl 2014;2:85–92.
[29] Jones B, Goos P. I-optimal versus D-optimal split-plot Response surface designs. J Qual
Technol 2012. doi:10.1080/00224065.2012.11917886.
[30] AL-Farraji AAM. Chemical Engineering and Reactor Design of a Fluidised Bed Gasifier
. Cardiff University, 2017.
[31] Shah YT. Energy and fuel systems integration. 1st Edition. CRC Press ; 2015.
[32] Moni MNZ, Sulaiman SA, Hassan S. Potentials of Selected Malaysian Biomasses as
Co-Gasification Fuels with Oil Palm Fronds in a Fixed-Bed Downdraft Gasifier.
MATEC Web Conf 2014;13:06004. doi:10.1051/matecconf/20141306004.
[33] Rajvanshi AK. Biomass gasification. Altern Energy Agric 2014;II:1–21. doi:0-8493-

32
ACCEPTED MANUSCRIPT

6348-9.
[34] ÖzyuǧUran A, Yaman S. Prediction of Calorific Value of Biomass from Proximate
Analysis. Energy Procedia, vol. 107, 2017, p. 130–6. doi:10.1016/j.egypro.2016.12.149.
[35] Nguyen TLT, Gheewala SH, Garivait S. Full chain energy analysis of fuel ethanol from
cane molasses in Thailand. Appl Energy 2008;85:722–34.
doi:10.1016/j.apenergy.2008.02.002.
[36] Czernik S, Bridgwater A V. Overview of Applications of Biomass Fast Pyrolysis Oil.
Energy & Fuels 2004;18:590–8. doi:10.1021/ef034067u.
[37] Asadullah M, Ab Rasid NS, Kadir SASA, Azdarpour A. Production and detailed
characterization of bio-oil from fast pyrolysis of palm kernel shell. Biomass and
Bioenergy 2013;59:316–24. doi:10.1016/j.biombioe.2013.08.037.
[38] Kuhe A, Iortyer HA, Kucha EI. Experimental investigation of a “throatless” downdraft
gasifier with palm kernel shell. J Renew Sustain Energy 2013;5. doi:10.1063/1.4811802.
[39] Shawal Nasri N, Jibril M, Ahmad Zaini MA, Mohsin R, Usman Dadum H, Murtala Musa
A. Synthesis and characterization of green porous carbons with large surface area by
two step chemical activation with KOH. J Teknol (Sciences Eng 2014;67:25–8.
doi:10.11113/jt.v67.2787.
[40] Daud WMAW, Ali WSW. Comparison on pore development of activated carbon
produced from palm shell and coconut shell. Bioresour Technol 2004;93:63–9.
doi:10.1016/j.biortech.2003.09.015.
[41] Tilay A, Azargohar R, Gerspacher R, Dalai A, Kozinski J. Gasification of Canola Meal
and Factors Affecting Gasification Process. Bioenergy Res 2014;7:1131–43.
doi:10.1007/s12155-014-9437-5.
[42] Çaglar A, Demirbaş A. Hydrogen-rich gaseous products from tea waste by pyrolysis.
Energy Sources 2001;23:739–46. doi:10.1080/00908310120370.
[43] Narváez I, Orío A, Aznar MP, Corella J. Biomass gasification with air in an atmospheric
bubbling fluidized bed. Effect of six operational variables on the quality of the produced
raw gas. Ind Eng Chem Res 1996;35:2110–20. doi:10.1021/ie9507540.
[44] Raheem A, Ji G, Memon A, Sivasangar S, Wang W, Zhao M, et al. Catalytic gasification
of algal biomass for hydrogen-rich gas production: Parametric optimization via central
composite design. Energy Convers Manag 2018. doi:10.1016/j.enconman.2017.12.041.
[45] Shahbaz M, Yusup S, Inayat A, Patrick DO, Pratama A, Ammar M. Optimization of
hydrogen and syngas production from PKS gasification by using coal bottom ash.
Bioresour Technol 2017;241:284–95. doi:10.1016/j.biortech.2017.05.119.

33
ACCEPTED MANUSCRIPT

[46] Inayat M, Sulaiman SA, Kurnia JC. Catalytic co-gasification of coconut shells and oil
palm fronds blends in the presence of cement, dolomite, and limestone: Parametric
optimization via Box Behnken Design. J Energy Inst 2018.
doi:10.1016/j.joei.2018.08.002.
[47] He M, Xiao B, Liu S, Guo X, Luo S, Xu Z, et al. Hydrogen-rich gas from catalytic steam
gasification of municipal solid waste (MSW): Influence of steam to MSW ratios and
weight hourly space velocity on gas production and composition. Int J Hydrogen Energy
2009;34:2174–83. doi:10.1016/j.ijhydene.2008.11.115.
[48] Lahijani P, Zainal ZA. Gasification of palm empty fruit bunch in a bubbling fluidized
bed: A performance and agglomeration study. Bioresour Technol 2011;102:2068–76.
doi:10.1016/j.biortech.2010.09.101.
[49] Ahmad AA, Zawawi NA, Kasim FH, Inayat A, Khasri A. Assessing the gasification
performance of biomass: a review on biomass gasification process conditions,
optimization and economic evaluation. Renew Sustain Energy Rev 2016;53:1333–47.
doi:10.1016/j.rser.2015.09.030.
[50] Madadian E, Orsat V, Lefsrud M. Comparative study of temperature impact on air
gasification of various types of biomass in a research-scale down-draft reactor. Energy
and Fuels 2017. doi:10.1021/acs.energyfuels.6b03489.
[51] Hernández JJ, Aranda-Almansa G, Serrano C. Co-gasification of biomass wastes and
coal-coke blends in an entrained flow gasifier: An experimental study. Energy and Fuels,
2010. doi:10.1021/ef901585f.
[52] Wongsiriamnuay T, Kannang N, Tippayawong N. Effect of operating conditions on
catalytic gasification of bamboo in a fluidized bed. Int J Chem Eng 2013.
doi:10.1155/2013/297941.
[53] Vélez JF, Chejne F, Valdés CF, Emery EJ, Londoño CA. Co-gasification of Colombian
coal and biomass in fluidized bed: An experimental study. Fuel 2009;88:424–30.
doi:10.1016/j.fuel.2008.10.018.
[54] Ariffin MA, Wan Mahmood WMF, Mohamed R, Mohd Nor MT. Performance of oil
palm kernel shell gasification using a medium-scale downdraft gasifier. Int J Green
Energy 2016;13:513–20. doi:10.1080/15435075.2014.966266.
[55] Xiao X, Meng X, Le DD, Takarada T. Two-stage steam gasification of waste biomass
in fluidized bed at low temperature: Parametric investigations and performance
optimization. Bioresour Technol 2011;102:1975–81.
doi:10.1016/j.biortech.2010.09.016.

34
ACCEPTED MANUSCRIPT

[56] Hernández JJ, Aranda G, San Miguel G, Bula A. Gasification of grapevine pruning
waste in an entrained-flow reactor: gas products, energy efficiency and gas conditioning
alternatives. Glob NEST J 2010;12:215–27.
[57] Wan Ab Karim Ghani WA, Moghadam RA, Salleh MAM, Alias AB. Air gasification of
agricultural waste in a fluidized bed gasifier: hydrogen production performance.
Energies 2009;2:258–68. doi:10.3390/en20200258.
[58] Carpenter DL, Bain RL, Davis RE, Dutta A, Feik CJ, Gaston KR, et al. Pilot-scale
gasification of corn stover, switchgrass, wheat straw, and wood: 1. Parametric study and
comparison with literature. Ind Eng Chem Res 2010. doi:10.1021/ie900595m.
[59] Abdoulmoumine N, Kulkarni A, Adhikari S. Effects of temperature and equivalence
ratio on mass balance and energy analysis in loblolly pine oxygen gasification. Energy
Sci Eng 2016. doi:10.1002/ese3.124.
[60] Dowaki K, Ohta T, Kasahara Y, Kameyama M, Sakawaki K, Mori S. An economic and
energy analysis on bio-hydrogen fuel using a gasification process. Renew Energy 2007.
doi:10.1016/j.renene.2005.12.010.
[61] Iwasaki W. A consideration of the economic efficiency of hydrogen production from
biomass. Int J Hydrogen Energy 2003. doi:10.1016/S0360-3199(02)00193-3.
[62] Mohammed MAA, Salmiaton A, Wan Azlina WAKG, Mohammad Amran MS,
Fakhru’L-Razi A, Taufiq-Yap YH. Hydrogen rich gas from oil palm biomass as a
potential source of renewable energy in Malaysia. Renew Sustain Energy Rev
2011;15:1258–70. doi:10.1016/j.rser.2010.10.003.
[63] Shahbaz M, Yusup S, Inayat A, Patrick DO, Pratama A, Ammar M. Optimization of
hydrogen and syngas production from PKS gasification by using coal bottom ash.
Bioresour Technol 2017. doi:10.1016/j.biortech.2017.05.119.
[64] Lv P, Wu C, Ma L, Yuan Z. A study on the economic efficiency of hydrogen production
from biomass residues in China. Renew Energy 2008.
doi:10.1016/j.renene.2007.11.002.

List of Figures

Fig. 1. Different particle sizes of (a) CS and (b) PKS with the average particle sizes of

8–11 mm (large), 4–7 mm (medium), and 1–3 mm (small).

35
ACCEPTED MANUSCRIPT

Fig. 2. Schematic representation of the batch-type downdraft fixed-bed reactor.

Fig. 3. 3D surface response of the gas composition for the CS: (a) H2 and (c) CO, and for the

PKS: (b) H2 and (d) CO.

Fig. 4. 3D surface response of the gas composition for the CS: (a) CO2 and (c) CH4, and for

the PKS: (b) CO2 and (d) CH4.

Fig. 5. 3D surface response of gasification performance for the CS: (a) HHVsyngas and (c) Y and

for the PKS: (b) HHVsyngas and (d) Y.

Fig. 6. 3D surface response of gasification performance for the CS: (a) c and (c)  and for the

PKS: (b) c and (d) 

List of Tables

Table 1. Ranges and levels of different process variables for the IV-optimal design.

Table 2. Characterization of the coconut shell (CS) and palm kernel shell (PKS).

Table 3. Overall material mass balance for gasification of the CS and the PKS.

Table 4. Overall carbon mass balance for gasification of the CS and the PKS.

Table 5. Operating conditions for gasification of the CS and PKS using a downdraft reactor.

Table 6. Cost analysis for the gasification of the CS and PKS using a downdraft reactor.

Table 7. Comparison of hydrogen production costs through different processes.

36
ACCEPTED MANUSCRIPT

Highlights

 Coconut and palm kernel shells were gasified in a fixed-bed downdraft reactor.

 Shell particle size and operating temperature affected the gasification performance.

 The quality of gas composition increased with a decrease in shell particle size.

 Temperature showed a significant effect on the gasification performance.

 High gasification performance was obtained with the coconut shells.

S-ar putea să vă placă și