You are on page 1of 15

Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Contents lists available at ScienceDirect

Renewable and Sustainable Energy Reviews


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

Assessing the gasification performance of biomass: A review


on biomass gasification process conditions, optimization
and economic evaluation
Anis Atikah Ahmad a,n, Norfadhila Abdullah Zawawi a, Farizul Hafiz Kasim a, Abrar Inayat b,
Azduwin Khasri c
a
School of Bioprocess Engineering, Universiti Malaysia Perlis, 02600 Jejawi Arau, Perlis, Malaysia
b
Department of Chemical Engineering, Faculty of Engineering, Universiti Teknologi Petronas, 31750 Tronoh, Perak, Malaysia
c
Department of Chemical Engineering Technology, Faculty of Engineering Technology, Universiti Malaysia Perlis, Kampus Uniciti Sg Chuchuh, 02100 Padang
Besar, Perlis, Malaysia

art ic l e i nf o a b s t r a c t

Article history: Currently, the use of biomass as an energy source has received a tremendous amount of interest from all
Received 17 March 2015 over the world due to its advantage in providing a continuous feedstock supply. Moreover, when com-
Received in revised form pared to fossil fuels, biomass fuels possess negligible sulfur concentrations, produce less ash, and gen-
24 May 2015
erate far less emissions into the air. Biomass has a potential to be a very promising alternative source of
Accepted 17 September 2015
Available online 10 November 2015
raw material for syngas production due to its tremendous availability. Syngas can be produced from the
gasification of a biomass. However, this process requires a significant amount of energy due to the
Keywords: endothermic behavior of the reaction. The energy consumption during the gasification process is a major
Gasification constraint on the thermal efficiency and on the design of the gasifier. Therefore, a substantial
Syngas
improvement and the optimization of the available gasification process are very crucial in developing a
Optimization
sustainable utilization of these renewable natural resources. This review article focuses on highlighting
Simulations
Economic evaluation the characteristics and performances of the different types of gasifiers under variable process parameters
that will affect the yields of the end products as well as the composition of the gas. The various types of
models used in the simulations, the optimization of the gasification conditions and the economic eva-
luation of the gasification process will also be discussed.
& 2015 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1334
2. Biomass gasification for the production of syngas. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1334
3. Gasification performance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335
3.1. Effect of the gasifier types. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1335
3.2. Effect of the temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1336
3.3. Effect of the biomass particle size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
3.4. Effect of the gasification agent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
3.5. Effect of the bed material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1341
4. Simulation of the gasification process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1343
4.1. Modeling of the biomass gasification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1344
5. Optimization of the gasification process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1345
6. Economic evaluation of the H2 gas production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1345

n
Corresponding author. Tel.: þ 60 12 5847 688; fax: þ 60 4 979 8755.
E-mail address: anisatikah@unimap.edu.my (A.A. Ahmad).

http://dx.doi.org/10.1016/j.rser.2015.09.030
1364-0321/& 2015 Elsevier Ltd. All rights reserved.
1334 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

7. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346


Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346

1. Introduction the biomass. The energy produced can be used to provide heat
and/or steam for cooking, space heating, and industrial processes,
Currently, the use of biomass as an environmentally friendly as well as for the generation of electricity. Biochemical processes
renewable energy source has received a tremendous amount of are processes that are composed of anaerobic digestion and alco-
interest from all over the world, as evidenced by the fact that the holic fermentation. Meanwhile, the example of the agrochemical
use of biomass has been estimated to contribute approximately process is the extraction method, which is mainly mechanical in
10–14% of the world's power supply [37]. Biomass possesses an nature, such as the extraction of rapeseed oil from rapeseed.
advantage over fossil fuels (natural gas, coal, and petroleum) in Thermochemical conversion processes can be further subdivided
that it provides a continuous feedstock supply. Moreover, the into gasification, pyrolysis, supercritical fluid extraction and direct
constant uncertainties of the supply levels and the pricing of fossil liquefaction. Thermochemical decomposition can be utilized for
fuels, as well as the greenhouse effect and its consequences on the energy conversion of all of the categories of biomass materials,
climate change, are some of the drawbacks resulting from the but low moisture, herbaceous (small grain field residues) and
extensive use of fossil fuels [43]. Biomass is a CO2 neutral resource woody (woody industry wastes and standing vegetation not sui-
for its life cycle [28,29] and it also possesses a zero CO2 net table for lumber) materials are the most suitable [13].
emission energy [38]. Compared to biomass, burning fossil fuels By definition, “Bio” means life and biomass is a biological
takes carbon that was locked away underground (as crude oil, gas, material with a large volume that was derived from living
and coal) and transfers it into the atmosphere as CO2. However, organisms, such as plant or animal waste [41]. Generally, biomass
biomass combustion, such as wood combustion, only recycles can be categorized into five basic categories. The first category is
carbon that was already in the natural carbon cycle, with the net the virgin wood that was obtained from forestry or in waste from
effect being that no new CO2 was added to the atmosphere as long forest products. Examples of this include palm kernel shells, wood
as the forests from which the wood came are sustainably managed pellets, woodchips, and sawdust. The second category can be
[58]. The comparison of the carbon cycles by using a fossil fuel and classified as the energy crops, which are the high-yield crops
a biomass feedstock is illustrated in Fig. 1. grown specifically for energy applications. Hybrid eucalyptus,
In terms of the environmental effects, biomass fuels possess a Jatropha, Pongamia, and perennial grasses such as Miscanthus and
negligible sulfur concentration, produce less ash, and generate far Napier grass growing on marginal land that is not fit for agri-
less air emissions in comparison with fossil fuels. Therefore, the culture are examples of this class. The third category is the agri-
combustion of a biomass does not contribute to sulfur dioxide cultural residues, which includes the bagasse from sugarcane, corn
emissions, which is the cause of acid rain, and the produced ash husks, coconut shells, and straw. The fourth category is the food
can be used as a soil additive for select farms. It should also be waste, which includes animal fat, residues from food and drink
noted that through the utilization of biomass as the energy source, manufacturing, etc. The fifth category is the industrial waste from
the amount of waste sent to landfills can be reduced, which would manufacturing and industrial processes [25]. Wood and wood
have an impact the significant waste disposal problem, particularly wastes are the major sources that contribute to the biomass
in municipal areas. The use of biomass as an energy source also energy (64%), followed by MSW (24%), agricultural waste (5%), and
possesses an economical advantage. The use of biomass as an landfill gases (5%) [13].
energy source would not be affected by the world price fluctua-
tions or the uncertainties in the supplies of imported fuels.
Therefore, a reduction on the dependency on fossil fuels, such as 2. Biomass gasification for the production of syngas
oil, would reduce the economic pressures of importing petroleum
products [13]. Syngas is a key intermediate in the chemical industry. It is used
Biomass conversion technologies can be categorized into four in a number of highly selective syntheses of a wide range of che-
basic categories: direct combustion processes, biochemical pro- micals and fuels, such as Fischer–Tropsch liquids, methanol and
cesses, agrochemical processes and thermochemical processes. ammonia [35]. It is also used as a source of pure hydrogen and
Direct combustion is the main process for obtaining energy from carbon monoxide [49]. Currently, hydrogen is the most promising

Fig. 1. Comparison of carbon flows for fossil fuel and biomass heating systems [58].
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1335

energy source that can be used in fuel cells and internal com- intermediate energy carriers that can either be combusted for heat
bustion engines [5]. Syngas can be produced from biomasses using and power generation or they can further be synthesized into
thermochemical processes, such as pyrolysis, gasification, and transportation fuel. The produced gas can also be used as a feed-
reforming or combustion processes. Meanwhile, the biological stock for the production of high-value chemicals, such as olefins
processes used to produce syngas are direct biophotolysis, indirect and formaldehyde [53]. Fig. 3 summarizes the steps involved in
biophotolysis, biological water–gas shift reactions, photo- the gasification process.
fermentation and dark-fermentation [43]. The production and composition of the production gas varies
Gasification is the most effective process for the production of widely with the properties of input streams, the employed gasi-
hydrogen from biomasses. Syngas and biomass gasification were fication reactor type, the operating conditions, and the gas output
known before World War II, and the first gasification plant was conditioning. Input streams are referred to by feedstock materials
established in North America in 2001 [54]. Gasification utilizes (type and components) and the gasifying agent or medium (type
various types of carbon-based feedstocks, such as natural gas, coal, and flow of gasifying agent). Gas output conditioning is composed
petroleum, petcoke, biomass, and industrial wastes [56]. As illu- of cooling and the disposal of particulate matter and tar in the gas
strated in Fig. 2, the chart shows that coal now dominates as the product [15,62]. The operating conditions are composed of tem-
feedstock in 51% or 36,315 MWth of syngas capacity, representing perature, pressure, and the residence time are used to control the
53 plants. Petroleum (including fuel oil, refinery residue, and gasification reactions [62]. Recently, there has been research con-
naphtha) is the second leading feedstock, with 17,938 MWth, or ducted on biomass gasification utilizing different operating con-
25% of total gasification capacity, representing 56 plants. Natural ditions, such as (1) the types of gasifiers, namely, fixed bed, flui-
gas provides 22%, petcoke provides approximately 1%, and the dized bed, and entrained flow [19]; (2) the gasification agents,
least used feedstock is biomass or waste, which accounts for such as air, oxygen, steam, CO2, or mixtures of these components
0.5% [56]. [43]; and (3) operating conditions that include the temperature,
Biomass gasification is the conversion of carbonaceous material pressure, catalyst and the equivalent that have an effect on the
into a gaseous product or synthesis gas that mainly consists of yields of the end products and the composition of the gas.
hydrogen (H2) and carbon monoxide (CO), with lower amounts of
carbon dioxide (CO2), water (H2O), methane (CH4), higher hydro-
carbons (C2 þ), and nitrogen (N2) [56]. The gasification process is 3. Gasification performance
performed in the presence of a gasifying agent (for example air,
pure oxygen, or steam, or mixtures of these components) at ele- There are a few parameters that have an effect on the perfor-
vated temperatures between 500 and 1400 °C and at atmospheric mance of the syngas production during gasification. The types of
or elevated pressures up to 33 bar (480 psia) [57]. According to Ref. the gasifier and the operating conditions (such as the temperature,
[43], the gasification process is applicable for biomasses that gasification agent, biomass particle size and type of bed materials)
possess a moisture content lower than 35%. For biomass feed- should be thoroughly investigated in order to design the best
stocks that possess higher amounts of moisture, in the range of gasification system. These factors will be further highlighted in
25–60%, the use of these feedstocks directly in the gasifier will following sections.
result in great amounts of energy losses in the overall process. It
was recommended that the biomasses be preheated or dried to 3.1. Effect of the gasifier types
moisture contents between 10% and 20% before they are intro-
duced into the gasifier [40]. The gasifiers can be divided into different types such as the
The first step in gasification is the devolatilization of the bio- fixed bed (updraft and downdraft), moving bed, fluidized bed
mass particles, which produces vapors and char, followed by the (bubbling and circulating), and the entrained flow gasifier. The
cracking and reforming of the volatiles and the gasification of the gasifiers are categorized into different types based on how the
char. A steam atmosphere enhances the reforming reactions and biomass is supported in the reactor vessel, the direction of flow of
char gasification so that more light gases, such as H2, CO and CO2, both the biomass and the oxidant, and the way the heat is supplied
can be produced [31]. The generated gas mixtures are to the reactor. In a fixed bed gasifier, the biomass bed is held

Fig. 2. World gasification operating capacity by feedstock [56].


1336 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Fig. 3. Gasification process flow and its applications.

stationary while the reaction front passes through it, or the bed the gasifying medium in the gasifier, large quantities of tar and
can move through the reaction or it can undergo a mechanical char were generated. Conversely, fluidized bed gasifiers allowed
displacement. Fixed bed gasifiers possess two common config- for good mixing and good gas–solid contact, which increased the
urations, one of which is a fixed bed updraft or counter current reaction rate and the conversion efficiencies. Furthermore, a lower
gasifier and the other is a fixed bed downdraft or cocurrent gasi- concentration of tar in the product gas was achieved with the
fier. The fixed bed updraft or counter current gasifier configuration usage of the bed material as the heat transfer medium and cata-
is the oldest and simplest form of a gasifier and is noted to have a lyst, and thus improving the quality of the gas. In addition, effi-
relatively low cost owing to the simple reactor concept. Biomass cient heat and mass transfer properties in the fluidized bed gasi-
feedstock is introduced at the top of the reactor and a gasification fier enabled the usage of assorted types of biomass wastes with
agent is introduced below the gasifier. The produced gas is different compositions and heating values [3,65].
extracted at the top of the gasifier. Although the fixed bed
downdraft (cocurrent gasifier) has the same mechanical config- 3.2. Effect of the temperature
uration as the updraft gasifier [56], both the oxidant and the
biomass are introduced at the top of these reactors [15]. According to Ref. [57], when temperatures greater than 1200–
The fluidized bed gasifiers are categorized based on their fluid 1300 °C were utilized, the following results were obtained: little to
dynamics and modes of heat transfer. There are two common no methane, the formation of higher hydrocarbons or tar, and a
types of fluidized bed gasifiers, which are the bubbling fluidized maximized production of H2 and CO without the need for further
bed and the circulating fluidized bed. A bubbling fluidized bed conversion steps [56]. Rapagnà and Latif [48], who studied the
usually consists of fine, inert particles of sand or alumina. As the steam gasification of almond shells in a continuous fluidized bed
gasification agent is forced through the inert particles, a point is reactor, reported that the highest yield of total gas, H2, CO, CH4,
reached when the frictional force between the particles and the and CO2 obtained were 1.55, 0.7, 0.42, 0.12, and 0.300 m3/kg bio-
gas counterbalances the weight of the solids. At this gas velocity mass, respectively. The study indicated that by increasing the
(minimum fluidization), a bubbling and channeling of the gas temperature from 600 to 800 °C, the concentration of char and
through the media occurs such that the particles remain in the heavy tars decreased while the overall gas yield increased, and the
reactor and appear to be in a “boiling state”. In contrast, circulating yields of H2 and CO also increased. Similarly to the studies of Refs.
fluidized bed gasifiers operate at gas velocities above the mini- [48,38] also performed a gasification in the fluidized bed using
mum fluidization velocity which results in the entrainment of the empty fruit bunch (EFB) as the feedstock and air as the gasifier
particles in the gas stream. The entrained particles in the gas exit agent in the temperature range between 700 and 1000 °C. When
the top of the gasifier, are separated in a cyclone, and are then the temperature increased from 700 to 1000 °C, the H2 con-
returned back to the gasifier [57]. Normally, smaller bed particles centration increased from 10.27 to 38.02 vol%, the CO2 con-
are used in the circulating fluidized bed [45]. Compared to the centration decreased, of the concentration of CO increased, and
bubbling fluidized bed, the circulating fluidized bed offers higher the concentration of CH4 increased from 5.84 to 14.72 vol%. By
conversion rates and efficiencies [15]. Based on a survey of the utilizing a higher temperature, the percentage of char and tar
present commercial or near-commercial biomass gasification decreased, the LHV value increased, and the total gas yield
technologies, directly heated bubbling fluidized bed (BFB) gasifiers increased then reached a maximum value of  92 wt% when the
are the most widely used gasifier that are operated with a broad temperature was 1000 °C. In other research, Lv et al. [32] varied
span of parameters, such as the temperature, pressure and the temperature in the fluidized bed reactor from 700 °C to 900 °C
throughput [56]. and used sawdust as the feedstock material. The results indicated
There are several studies that compare the advantages and that higher temperatures contributed to higher gas yields and
disadvantages of the fixed bed and the fluidized bed gasifier. increased hydrogen production, while the concentration of CO,
Warnecke [61] compared the fixed bed and fluidized bed using CH4, and CO2 decreased. The gas yield and hydrogen composition
different criteria, such as the technology, use of material, use of increased from 1.43 Nm3/kg biomass to 2.53 Nm3/kg biomass and
energy, and the environmental and economic impact. From their 21 vol% to 39 vol%, respectively, with the increase in temperature.
research, Warnecke [61] concluded that the advantages of one of The carbon conversion efficiency (CCE) also increased with the
the reactor types over the other were marginal and that there was temperature increase. When the temperature was increased from
no significant advantage between the fixed bed and the fluidized 700 to 750 °C, the LHV value first decreased, and increased to its
bed reactor. Meanwhile, according to Ref. [3], there are several maximum value of 8.56 MJ/Nm3 when the temperature was fur-
drawbacks to fixed bed gasifiers. Due to the low and non-uniform ther increased to 800 °C. The LHV value started to decrease when
heat and mass transfer between the solid biomass feedstock and the temperature was further increased from 800 to 900 °C, and the
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1337

lowest LHV value was recorded at 900 °C. Li et al. [27] studied the yields remained almost constant (32.99 and 33.93 vol%) for par-
steam gasification of EFB in a fixed bed reactor using a tri-metallic ticles sizes of o0.3 mm and between 0.3 and 0.5 mm, respec-
catalyst in the temperature range between 750 and 900 °C, and tively, and then they decreased to 21.57 vol%. For particle sizes
the results demonstrated that higher temperatures contributed to between 0.5 and 1.0 mm. The highest LHV of the product gas and
increased hydrogen production and total gas yields. The maximum the optimum product gas composition were achieved when the
value of the total gas and hydrogen gas yields were 2.48 m3/kg and feedstock particle size was in the range of 0.3–0.5 mm. Similarly, Li
1.481 m3/kg, respectively, which was obtained at temperature of et al. [27] reported that by using particle sizes between o 0.15 and
900 °C. As the temperature increased from 750 to 900 °C, the 5 mm, the total gas yield and H2 yield decreased. The smaller
concentrations of H2 and CO2 increased, while the concentrations particles also produced more H2 and CO2 and less CH4, CO and
of CO and CH4 decreased. The LHV of the product gas decreased C2H4 than the larger particles. However, the LHV value increased
from 11.26 to 9.13 MJ/Nm3 when temperature was increased from with an increase in particle size, and a maximum value of
750 to 900 °C. However, the gasification process of bamboo in a 10.28 MJ/Nm3 was obtained using the largest particle size. In
fluidized bed reactor using air in a study that was conducted by another research conducted by Ref. [32], the smaller particles
Ref. [62] yielded a different trend with regards to the effect of produced more CH4,CO and less CO2. The smallest particles also
temperature. With an increase in temperature, the concentration produced the highest gas yields, the highest LHV and the highest
of H2 and CO decreased, while the concentration of CO2 increased. carbon conversion efficiency, which were 1.53–2.57 Nm3/kg bio-
When the temperature was increased from 400 °C to 600 °C, the mass, 8.7 MJ/Nm3, and 95.10%, respectively. In the research con-
total gas yield remained constant. The maximum LHV value and ducted by Ref. [31], the biomass particle sizes between o0.075
the CCE were recorded when the operating temperature was and 1.2 mm were used in the steam gasification of pine sawdust at
500 °C. Table 1, which can be found below, summarizes several of a temperature of 900 °C. In this study, the total gas yield, H2 gas
the investigations on the effect of temperature on the gasification yield and carbon conversion efficiency decreased while the con-
of various biomasses. centration of char and tar increased when the particle size
From these studies, it can be concluded that, the higher tem- increased. The highest overall gas yield and H2 gas yield obtained
peratures contributed to lower concentrations of char and heavy were 1.62 m3/kg biomass and 0.8 m3/kg biomass, respectively,
tars and higher concentrations of H2 gas as well as an increase in when the smallest particle size was used. The study also deter-
the overall gas yield. Higher temperature contributed to higher gas mined that the smaller particles produced more CO and less CO2
yield due to release of more volatiles. The increase in H2 produc- than the larger particles. However, a different trend was observed
tion was due to the tar thermal cracking reaction, which also in the study conducted by Ref. [64] using a biomass feedstock
decrease the tar concentration [51]. According to Le Chatelier's particle size from o 0.15 mm to 3 mm. At a temperature of 850 °C,
principle, increased temperature favors the products of endo- when particle size was increased from o0.15 mm to 0.45–0.9 mm,
thermic reactions and favors the reactants in exothermic reactions. the dry gas yield first gradually increased from 1.72 to 1.84 Nm3/
Therefore, the endothermic reforming reaction of hydrocarbon kg, but then it slightly declined to 1.81 Nm3/kg as the particle size
were improved with the increasing temperature (boudouard (R2), increased. A maximum dry gas yield of 1.84 Nm3/kg was obtained
water–gas (R3), and steam-methane reforming (R8) in Table 6). at a particle size of 0.45–0.9 mm. Meanwhile, the composition of
Increase in water–gas (R3) and steam-methane reforming (R8) the product gas remained nearly constant when the particle size
reactions resulted in increased H2 concentration [27]. The increase was varied, which demonstrated that the particle size had no
activity of steam-methane reforming (R8) also resulted in a effect on the molar fraction of syngas. Several studies on the effect
decrease of CH4 concentration and an increase in H2 and CO [32]. of the biomass particle size are summarized in Table 2.
CO production was favored with increased temperature due to Smaller particle sizes contributed to higher total gas yields,
boudouard (R2), water–gas (R3), and steam-methane reforming higher H2 concentrations, and lower char and tar yields. Larger
(R8) [51]. Meanwhile, the exothermic reaction of char partial feedstock particle size increase the temperature gradient inside
combustion (R1) affect the CO composition in the producer gas. the particle, so that at a given time the core of the particle has
Higher temperature was not favorable for CO production, so the lower temperature compared to the particle surface, which
content of CO decreased with temperature. The increase of CCE resulted to the increase of the char and liquids yields and decrease
was contributed due to more carbon and steam conversion in gases [38]. Smaller particles can obtain faster heating rate due
through endothermic reaction of boudouard reaction (R2) and to larger surface area. High heating rates produce more light gases
water–gas reaction (R3) [32]. Generally, higher temperature and less char and condensate. Besides that, the smaller particle
favored hydrogen production and gas yield but did not always size has greater contact area of biomass and steam, which resulted
favor gas heating value as too high a temperature lowered gas to higher chemical reaction rates and sufficient gasification reac-
heating value [32]. tions [27]. Several studies concluded that the effect of particle size
was less important compared to other parameters. Although larger
3.3. Effect of the biomass particle size particle size increased reaction time needed for the completion of
reactions, but at higher temperature, the effect of particle size on
In addition to the effect of temperature, some researchers have gasification may decreased.
also determined that the feedstock particle size has an effect on
the gasification product. From the research conducted by Ref. [38], 3.4. Effect of the gasification agent
a biomass particle size between o0.3 and 1.0 mm was studied,
and it was determined that feedstock particle sizes not in this Currently, there are many studies of biomass gasification using
range generally caused a blockage in the feeder. In this study, the different agents such as air, oxygen, steam, or mixtures of these
total gas yield decreased and the char and tar yields increased components. Several studies on the effect of gasification agent are
with an increase in the particle size. The highest gas yield that was summarized in Table 3. Lv et al. [34] made a comparison between
obtained was 74.79 wt% using the smallest particle size air and steam as a gasification agent and reported that gasification
(o 0.3 mm), and the lowest gas yield of 72.74 wt% was obtained using steam was more effective in the maximizing the yield of
using the largest particle size (0.5–1.0 mm). The study also deter- hydrogen compared to air. 0.49 m3 and 0.33 m3 of hydrogen per kg
mined that smaller particle sizes produced more CO and CH4 and of biomass were produced by gasifying pine wood in a downdraft
lower amounts of CO2 than the larger particles; the hydrogen gasifier (fixed-bed) with steam and air, respectively. It was observed
1338 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Table 1
Effect of the temperature.

No. System configuration Gas yield Gas composition Biomass remaining CCE (%) Gas LHV Conditions for the optimum Reference
and operation as char and heavy result
parameters tars

1 GT: continuous flui- Gas yield: (m3/kg na 0–35 na na BT: 800 °C and FS: 0.287 mm [48]
dized bed biomass) (maximum total gas yield and H2
ID: 60 mm Total gas: 0.5– gas yield, and low char and tars)
F: almond shells 1.55
FR: 0.06 kg/h H2: 0.175–0.7
FS: 0.287, 0.533, 0.75,
1.09 mm
BM: fine alumina CO: 0.125–0.42
GA: steam
SB: 0.8
BT: 600, 650, 700, 750, CH4: 0.050–0.12
800 °C CO2: 0.125–0.3
2 GT: fluidized bed Total gas: 62.68– H2: 10.27–38.02 vol% 6–18 wt% na 7.5– BT: 1000 °C (maximum total gas [38]
L: 600 mm 91.7 wt% 15.55 MJ/m3 yield, optimum gas composition,
ID: 40 mm CO: 21.87–36.36 vol highest LHV, and low char and
F: EFB (o 10 wt% % tars)
moisture content) CH4: 5.84–14.72 vol
FR: 0.6 kg/h %
FS: 0.3–1 mm CO2: 10–65 vol%
BM: inert sand
GA: air
ER: 0.15–0.35
BT: 700, 800, 900,
1000 °C
3 GT: fluidized bed Total gas: 1.43– H2: 21–39 vol% na 78.17- 7.362– BT: 900 °C (maximum gas yield, [32]
H: 1400 mm 2.53 Nm3/kg 92.59 8.56 MJ/Nm3 optimum gas composition, and
ID: 40 mm biomass CO: 35–43 vol% highest CCE)
F: pine sawdust CH4: 6–10 vol%
FR: 0.445 kg/h
FS: 0.3–0.45 mm CO2: 18–20 vol% BT: 800 °C (highest LHV)
BM: silica sand
GA: air–steam
SB: 2.7
ER: 0.22
BT: 700, 750, 800,
850, 900 °C
4 GT: fixed bed Total gas: 1.79– H2: 48–60 vol% na na 9.13– BT: 900 °C (maximum total gas [27]
ID: 88 mm (catalytic 2.48 m3/kg CO: 15–25 vol% 11.26 MJ/ yield and H2 gas yield, optimum
reactor) Nm3 gas composition)
H: 1200 mm (catalytic H2: 0.861– CH4: 5–5 vol%
reactor) 1.481 m3/kg
F: palm oil wastes CO2: 20–25 vol%
(shell, fiber, and EFB)
FR: 0.3–1 kg/h C2 hydrocarbons: 1– BT: 750 °C (highest LHV)
FS: 0.15-2 mm 2 vol%
BM: tri-metallic
catalyst
GA: steam
SB: 1.33
BT: 750, 800, 850,
900 °C
5 GT: fluidized bed Total gas: 1.9– H2: 6.6–8.16 mol% na 63.6– 1.6–1.9 MJ/ T: 400 °C (highest H2 content) [62]
H: 2000 mm 2.0 Nm3/kg 67.4 m3
ID: 50 mm biomass CO: 23.5–30.6 mol%
F: bamboo T: 500 °C (highest CCE and LHV)
FR: 0.6 kg/h CH4: 4–5 mol%
FS: 0.10–0.25 mm CO2: 59-63 mol%
BM: silica sand
GA: air, ER: 0.4
BT: 400, 500, 600 °C

GT: gasifier types, L: gasifier length, H: gasifier height, ID: gasifier internal diameter, BM: bed material, GA: gasifying agent, BT: bed temperature, F: feedstock, FR: feed rate,
FS: feedstock size, ER: equivalence ratio, SB: steam to biomass ratio, CCE: carbon conversion efficiency, na: data were not available.
Summary: Higher temperatures contributed to lower concentrations of char and heavy tars and higher concentrations of H2 gas as well as an increase in the overall gas yield.

that the concentration of H2 and CO reached 63.27–72.56% for the instead of using just air. The H2 and CO gas compositions in the
biomass oxygen/steam gasification, while the concentration of H2 product gas were also higher when an air–steam mixture was used
and CO only reached 52.19–63.31% for biomass air gasification. The as the gasifying medium. In contrast, Gil et al. [18] reported that a
same trend was observed in the research conducted by Ref. [62]. In gasification with air produced a gas yield of 1:4  2:4 m3n dry basis/
their research, the value of the CCE and the total gas yield increased kg biomass daf that was significantly larger than what was pro-
when air–steam was used as the medium in the gasification process duced in the gasification with steam, which was 0:8 1:1 m3n dry
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1339

Table 2
Effect of biomass particle size.

No. System configuration and Gas yield Gas composition Biomass remain- CCE (%) Gas LHV Conditions for the optimum result Reference
operation parameters ing as char and
heavy tars

1 GT: fluidized bed Total gas: H2: 21.57–33.93 vol% 13–18 wt% na 11.8– FS: o0.3 mm (maximum gas [38]
L: 600 mm 72.74–74.79 wt 15.26 MJ/m3 yield, low char and heavy tar)
ID: 40 mm % CO: 35–42.5 vol%
F: EFB ( o10 wt% moist- CH4: 15–17.5 vol% FS: 0.3–0.5 mm (optimum gas
ure content) composition and highest LHV)
FR: 0.6 kg/h CO2: 7.5–30 vol%
FS: o 0.3,0.3–0.5,0.5–
1.0 mm
BM: inert sand
GA: air,
ER: 0.15–0.35
BT: 850 °C
2 GT: fixed bed Total gas: 2.16– H2: 55–58 vol% na na 8.99– FS: o0.15 (maximum gas yield, [27]
ID: 200 mm; 88 mm (cat- 2.41 m3/kg CO: 14–18 vol% 10.28 MJ/ maximum H2 yield and optimum
alytic reactor) Nm3 gas composition)
H: 400 mm; 1200 mm H2: 1.183– CH4: 3–5 vol%
(catalytic reactor) 1.4 m3/kg
F: palm oil wastes CO2: 20–23 vol%
FR: 0.3–1 kg/h FS: 2–5 mm (Highest LHV value)
FS: o 0.15, 0.15-1, 1–2, C2 hydrocarbons:
and 2–5 mm 1–2 vol%
BM: tri-metallic catalyst
GA: steam,
SB: 1.33
BT: 800 °C; 850 °C (cata-
lytic reactor)
3 GT: fluidized bed Total gas: 1.53- H2: 30–32 vol% na 77.62– 7.0–8.7 MJ/ FS: 0.2–0.3 mm (maximum gas [32]
H: 1400 mm 2.57 Nm3/kg 95.10 Nm3 yield, optimum gas composition,
ID: 40 mm biomass CO: 16–20 vol% highest LHV, highest CCE)
F: pine sawdust
FR: 0.512 kg/h CH4: 6–7 vol%
FS: 0.2–0.3, 0.3–0.45, CO2: 16–20 vol%
0.45–0.6, and 0.6–0.9 mm
BM: silica sand
GA: air–steam
SB: 1.56; ER: 0.23
BT: 800 °C
4 GT: fixed bed Total gas: 1.38- H2: 40–51.2 vol% 0.4–10% 80– na FS: o0.075 (maximum total gas [31]
H: 600 mm 1.62 m3/kg 99.87 and H2 gas yield, optimum gas
OD: 219 mm biomass CO: 15–22.4 vol% composition, highest CCE, and low
F: pine sawdust char and tar)
FR: 0.3 kg/h H2: 0.55-0.8 m3/ CH4: 2–5 vol%
FS: o 0.075, 0.075–0.15, kg biomass CO2: 12–40 vol%
0.15–0.3, 0.3–0.6, 0.6–
1.2 mm
BM: calcined dolomite
GA: steam
ER: 1.2
BT: 900 °C
5 GT: fixed bed 1.72-1.84 Nm3/ H2: 47–49 wt% na na na FS: 0.45–0.9 mm (maximum total [64]
H: 1000 mm kg biomass CO: 14–15 vol% gas yield,)
ID: 50 mm
F: Char derived from CH4: 2 vol%
cyanobacterial blooms CO2: 30–35 vol%
(CDCB)
FS: o 0.15 mm, 0.15–
0.3, 0.3–0.45, 0.45–0.9,
0.9–3 mm
GA: steam
BT: 850 °C

GT: gasifier types, L: gasifier length, H: gasifier height, ID: gasifier internal diameter, BM: bed material, GA: gasifying agent, BT: bed temperature, F: feedstock, FR: feed rate,
FS: feedstock size, ER: equivalence ratio, SB: steam to biomass ratio, CCE: carbon conversion efficiency, na: data were not available.
Summary: Lower particle sizes contributed to higher total gas yields, higher H2 concentrations, and lower char and tar yields.

basis/kg biomass daf. The steam/biomass ratio (SB) also affected the decrease. The same trend was observed with the H2 gas composi-
gasification output. In the research conducted by Ref. [27], the tion in the product gas. The percentage of the H2 gas composition
increase of the SB from 0 to 1.33 resulted in an increase of the total increased significantly when the SB was increased from 0 to 1.33,
gas and hydrogen yield to their highest values of 2.48 m3/kg and but it slightly decreased when the SB ratio was further increased
1.481 m3/kg, respectively. However, when the SB was increased from 1.33 to 2.67. These results were in agreement with the results
from 1.33 to 2.67, the total gas and hydrogen yield started to obtained by Ref. [10], who studied the effects of α-cellulose
1340 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Table 3
Effect of the gasification agent.

No. System configuration Gas yield Gas composition Biomass remain- CCE (%) Gas LHV Conditions for the opti- Reference
and operation ing as char and mum result
parameters heavy tars

1 GT: fixed bed Using air; H2 and CO (using na na na GA: oxygen-steam [34]
(downdraft) Total gas: 0.82– air):
ID: 60 mm 0.94 Nm3/kg 52.19–63.31 vol%
H: 350 mm
F: pine wood blocks H2: 0.24–0.33 m3/kg H2 and CO (using
GA: Air, oxygen- Using O2/steam; oxygen/steam):
steam Total gas: 1.24- 63.27–72.56 vol%
1.62 Nm3/kg
H2: 0.36-0.49 m3/kg
2 GT: fluidized bed Total gas (using air): 1.9– H2 (using air): 6.6– na Using air: na GA: air–steam [62]
H: 2000 mm 2.0 Nm3/kg 8.16 mol% 63.6–67.4%
ID: 50 mm SB: & 1:1
F: bamboo CO (using air): 23.5– Using air– ER: 0.4
FS: 0.10–0.25 mm Total gas (using air– 30.6 mol% steam:
FR: 0.6 kg/hr steam): 2.8–2.9 Nm3/ 87.3–98.5%
BM: silica sand kg) H2 (using air–
GA: air & air– steam): 10.9–
steam 16.5 mol%
SB: 0:1 and 1:1 CO (using air–
ER: 0.4 steam): 36.1–
BT: 400, 500, 600 °C 40.3 mol%
3 GT: bubbling flui- Total gas (using air): 1:4 na na na na GA: air [18]
dized bed –2:4 m3n dry basis/kg
GA: air, pure steam biomass daf
Total gas (using
steam): 0:8–1:1 m3n
dry basis/kg biomass
daf.
4 GT: fixed bed Total gas: 1.2–2.48 m3/ H2: 47–58 vol% na na 8.73– SB: 1.33 (maximum gas [27]
ID: 200 mm; 88 mm kg CO: 14–33 vol% 11.98 MJ/ yield, maximum H2 yield,
(catalytic reactor) Nm3 and optimum gas
H: 400 mm; H2: 0.558–1.481 m3/kg CH4: 3–6 vol% composition)
1200 mm (catalytic CO2: 14–26 vol%
reactor)
F: palm oil wastes C2 hydrocarbons: 1– SB: 0 (highest LHV)
(shell, fiber, and EFB) 2 vol%
FR: 0.3 kg/h
FS: 0.15–2 mm
BM: tri-metallic
catalyst
GA: steam
SB: 0, 0.67, 1.33, 2,
2.67
BT: 800 °C
5 GT: fluidized bed Total gas: 0.78–1.02 m3/ H2: 13.5–18.56 vol% na na 6.55–7.61 SB: 1 (maximum gas yield, [10]
ID: 63.9 mm kg (MJ/Nm3) optimum gas composition)
H: 1100 mm CO: 6.45–11.21 vol%
F: α-cellulose CH4: 2.21–3.73 vol SB: 0 (highest LHV)
(moisture content %
2–10%) CO2: 26.3–27.77 vol
FS: o0.35 mm %
GA: air–steam
ER: 0.27
SB: 0, 0.5, 1, 1.5
BT: 800 °C
6 GT: fluidized bed Total gas: 70.75– H2: 18.37–27.42 vol% Char: 2.12– na 12.35– ER: 0.35 (maximum gas [38]
L: 600 mm 86.46 wt% 13.65 wt% 15.38 MJ/ yield)
ID: 40 mm CO: 32–45 vol% Tar:2.82– m3
F: EFB (o 10 wt% CH4: 12–15 vol% 9.83 wt% ER: 0.25 (maximum
moisture content) hydrogen content)
FR: 0.6 kg/h CO2: 16.66–
FS: 0.3–0.5 mm 36.05 vol% ER: 0.15 (highest LHV)
BM: inert sand
GA: air
ER: 0.15, 0.20, 0.25,
0.30, 0.35
BT: 850
7 GT: downdraft fixed Total gas: 0.6–0.8 wt% H2: 20–30 wt% 0.59–0.7 wt% na 8.8–10.4 MJ/ ER: 0.21 (maximum value [52]
bed reactor CO: 15–20 wt% m3 of gas yield, H2 content,
H: 500 mm and LHV)
ID: 12.5 mm CH4: 10–12 wt%
F: olive kernel
GA: air CO2: 40–55 wt%
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1341

Table 3 (continued )

No. System configuration Gas yield Gas composition Biomass remain- CCE (%) Gas LHV Conditions for the opti- Reference
and operation ing as char and mum result
parameters heavy tars

ER: 0.14, 0.21, 0.42


BT: 950

GT: gasifier types, L: gasifier length, H: gasifier height, ID: gasifier internal diameter, BM: bed material, GA: gasifying agent, BT: bed temperature, F: feedstock, FR: feed rate,
FS: feedstock size, ER: equivalence ratio, SB: steam to biomass ratio, CCE: carbon conversion efficiency, na: data were not available.
Summary: The usage of steam as gasification agent was more effective in maximizing the yield of hydrogen when compared to air. The yield of hydrogen can be enhanced
with the presence of steam. Nonetheless, an excess amount of steam present in the gasifier can cause the system to lose a lot of energy in order to heat the steam, which is
not favorable for the production of energy. Too high of value of the ER can cause the gas quality to be less attractive due to an increase in the oxidation reactions that lead to
an increase in the production of CO2 and a decrease in the production of combustible gases.

gasification at 800 °C and an equivalent ratio of 0.27 using SBs of 0, CO2 production and decrease in the production of combustible
0.5, 1, and 1.5. The H2 gas composition increased significantly from gases. Too high of an ER can result in lower concentrations of H2
13.50 vol% to 18.56 vol% when the SB was increased from 0 to 1. (R6) and CO with an increase in the CO2 concentration in the
However, it started to decrease when the SB ratio was further output gas (R5) [38]. However, higher equivalence ratios also lead
increased from 1 to 1.5. The SB value of 1 also produced the highest to exothermic oxidation reactions that, at the same time, offer
value of the total gas yield. The lowest LHV value was obtained more heat to the gasification process, which can optimize the
using an SB of 1.5. Meanwhile, for an air gasification process, the product quality to some extent (the destruction of tar) [52]. But,
equivalence ratio (ER) may affect the quality of the syngas that is when the ER value is too small, temperature in the gasifier is low,
produced. In the research conducted by Ref. [38], the ER was which is unfavorable for further reactions of biomass gasification
increased from 0.15 to 0.35, and as a result, the char and tar yields gas [34]. Generally, ER value is an important parameter to deter-
were reduced from 13.65 to 2.12 wt% and 9.83 to 2.82 wt%, mine the quality of gas produced from biomass gasification [34].
respectively. The product gas LHV decreased slightly from 15.38 to
12.35 MJ/m3. The total gas yield increased and recorded its highest 3.5. Effect of the bed material
value of 86.46 wt% at an ER of 0.35; the H2 concentration increased
to its maximum value of 27.42 vol% at an ER of 0.25 and started to There are several studies that have used sorbents to absorb the
decrease when the ER was increased from 0.25 to 0.35. For other CO2 gas that was produced from the gasification process. In the
gases, the CO2 gas increased significantly while the yields of CH4 experiment conducted by Ref. [36], the pine bark was gasified
and CO gas decreased. In another research conducted by Ref. [51], using a calcium oxide to biomass molar ratio (CaO/B) of one, which
when the ER was increased from 0.14 to 0.42, the concentration of was determined theoretically using thermodynamic calculations.
char and tar slightly decreased. Meanwhile, the total gas yield The total gas yield, H2 gas composition, carbon conversion effi-
increased to a maximum value of 0.8 wt% at an ER of 0.21, but it ciency and CO2 gas composition were 0.87 m3/kg, 60 vol%, 30.3%,
slightly decreased to 0.79 wt% at an ER of 0.42. Both the highest and 27.7 vol%, respectively, in the absence of the sorbent at a
value of the LHV and H2 gas composition was recorded at an ER of gasification temperature of 600 °C. Using the same parameters but
0.21, and the lowest values were recorded at an ER of 0.42. With an with the presence of sorbent, the total gas yield, H2 gas compo-
increase in the value of the ER, the CH4 and CO gas yields decreased, sition, and carbon conversion efficiency increased and the CO2 gas
but the CO2 gas yield increased and reached a maximum value at an composition decreased, and the corresponding values were
ER of 0.42. 1.42 m3/kg, 64.5 vol%, 55.6%, and 26.8 vol%, respectively. From
According to Ref. [10], the yield of hydrogen could be enhanced these results, it was concluded that the calcium oxide played a
with the presence of steam. Based from these studies findings, dual role of the sorbent and the catalyst. In another study,
higher SB ratio generally contributed to higher total gas yields, H2 Hanaoka et al. [20] also used calcium oxide as the CO2 sorbent in
yield, and H2 composition. Higher SB ratio enhanced the steam the steam gasification of Japanese oak. However, in contrast with
concentration, thus improve the steam reforming reactions (R3, R7, Ref. [36], Hanaoka et al. [20] studied the effects of the molar ratio
and R8), which resulted in the increase of gas yield, hydrogen yield of CaO to the carbon in the biomass, ([Ca]/[C]). In the absence of
and H2 and CO2 composition [27], but reduced the CO and CH4 CaO, the product gas contained CO2. Conversely, in the presence of
composition [22]. This trend agrees well with Le Chatelier's prin- CaO, ([Ca]/[C] ¼1, 2, and 4), no CO2 was detected in the product
ciple, which stated that, the reaction will shift to the side that would gas. The H2 yield obtained using [Ca]/[C] ¼0, 1, 2, and 4 were 0.5,
counteract the change. Increasing SB ratio influences the reaction to 0.7, 0.8, and 0.55 m3kg, respectively. At a [Ca]/[C] of 2, the max-
favor a direction that reduces the concentration of steam. Thus, the imum yield of H2 was obtained. One of the major issues in biomass
high steam rate promotes the gasification and methane reforming gasification is the problem of the formation of tar during the
forward to produce more H2 and also pushes the water gas shift process. Tar is a complex mixture of condensable hydrogen com-
reaction forward to give an upward trend of CO2. pounds that are composed of single ring to up to five ring aromatic
Nonetheless, excessive steam quantity in the gasifier can cause compounds along with other oxygen-containing hydrocarbons
the system to lose a large amount of energy in order to heat up the and complex poly aromatic hydrocarbons PAH [9]. Several studies
steam, which is not favorable for the production of energy [10]. An of the gasification also utilized catalyst particles for the purpose of
excess amount of steam present in the gasifier also may lowered improving the hydrogen yield and the decomposition of tar [50].
reaction temperature and then produced the low gas quality [27]. According to Ref. [50], the most popular bed particle is sand, which
Based from these studies findings, higher ER generally contributed performs very well mechanically, as evidenced by its wide
to lower char and tar yields, lower LHV, lower concentration of CO, industrial use in bubbling and circulating fluidized bed combus-
and higher concentration of CO2. According to Ref. [52], the ER tion applications. However, sand does not play any active role in
affects the gasification behavior in two opposite trends. As the ER biomass gasification. Conversely, a Ni catalyst, limestone, an
increases, the gas quality becomes less attractive due to an orthosilicate of iron and magnesium, etc., can meet the require-
increase in the oxidation reactions that lead to an increase in the ments of the mechanical activity and resistance [50]. According to
1342 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Table 4
Effect of the bed material.

No. System configuration and Gas yield Gas composition Biomass remain- CCE (%) Gas LHV Conditions for the optimum Reference
operation parameters ing as char and result
heavy tars (wt%)

1 GT: batch-type reactor Without Without CaO: na Without na Using CaO [36]
CaO: H2, CH4,CO, CO2 (vol CaO: 30.3
F: pine bark Total gas: %): 60, 3.2, 9.1, 27.7 Using
BM: CaO reagent 0.87 m3/kg CaO: 55.6
as CO2sorbent CaO/B:1 Using CaO:
Total gas: Using CaO:
GA: steam 1.42 m3/kg H2, CH4, CO, CO2 (vol
BT: 600 °C %): 64.5, 2.8, 5.9, 26.8
2 GT: batch reactor (autoclave) H2: 0.5– na na na na [Ca]/[C]: 2 (highest H2 yield) [20]
F: Japanese oak 0.8 m3/kg
BM: Ca(OH)2 powder as
CO2sorbent
[Ca]/[C]: 0, 1, 2, 4
GA: steam
BT: 650 °C
3 GT: fixed bed Total gas: H2: 55.4–68.3 mol% Char: 4.4–6.4 na na Catalyst loading: 0.8 wt% [59]
ID: 10.5 mm 0.91–1.3 m3/ (highest H2 content and low-
L: 500 mm kg CO: 20.2–36.9 mol% est char content)
F: glycerol
FS: 0.20–0.35 mm CH4: 2.4–5.9 mol%
BM: silicon carbide and CO2: 1.9–7.7 mol%
Ni/Al2O3catalyst
Catalyst loading: 0–0.8 wt%
GA: steam
BT: 800 °C
4 GT: fluidized bed Total gas: H2: 3.2–9.1 mol% na CCE: 59.5– na C/B: 1.5:1 (maximum H2 con- [62]
H: 2000 mm 1.9–2.1 Nm3/ 80.1 tent, Highest LHV, and highest
ID: 50 mm kg CO: 21.4–31.7 mol% CCE)
F: bamboo
FS: between 0.10 and
0.25 mm
FR: 0.6 kg/h
BM: silica sand and cal-
cined dolomite as catalyst
CB: (0:1, 1:1, and 1.5:1)
GA: air
ER: 0.4
T: 400–600 °C
5 GT: fixed bed Total gas: H2: 36.5–53.6 vol% Tar yield: 0.28– na 10.20– BM: newly developed tri- [27]
ID: 88 mm (catalytic 1.21– CO: 12.7–25.8 vol% 37.8 g/Nm3 12.72 MJ/ metallic catalyst (maximum
reactor) 2.11 m3/kg Nm3 gas yield, maximum H2 yield,
H: 1200 mm (catalytic H2: 0.442– CH4: 4.4–10.2 vol% and optimum gas
reactor) 1.131 m3/kg composition)
F: palm oil wastes (shell, CO2: 20.9–26.6 vol%
fiber, and EFB) C2 hydrocarbons:
FR: 0.3–1 kg/h 0.4–2.4 vol% BM: calcined dolomite
FS: 0.15–2 mm (highest LHV)
BM: no catalyst, calci-
neddolomite,nano-
NiLaFe/γ-Al2O3
GA: steam;
SB: 1.33
BT: 800 °C

GT: gasifier types, L: gasifier length, H: gasifier height, ID: gasifier internal diameter, BM: bed material, GA: gasifying agent, BT: bed temperature, F: feedstock, FR: feed rate,
FS: feedstock size, ER: equivalence ratio, SB: steam to biomass ratio, CCE: carbon conversion efficiency, CB: catalyst to biomass ratio, [Ca]/[C]: CaO to carbon in the biomass
molar ratio, na: data were not available.
Summary: Sand does not play an active role in biomass gasification. Calcium oxide played a dual role of the sorbent and the catalyst,and the total gas yield, H2 gas
composition, and carbon conversion efficiency increased with the presence of calcium oxide. Various catalysts were used as bed materials, such as a Ni-based catalyst and
calcined dolomite, which enhanced the tar reforming reaction and which decreased the tar yield and increased the total gas yield.

Ref. [63], Ni-based catalysts have been reported as one of the most biomass ratios (CB) of 0:1, 1:1, and 1.5:1 were studied. With an
promising catalysts, and they are effective and comparatively increase in the temperature and the CB, the yields of H2 and CO
cheaper compared with other noble metal catalysts. In the increased while the yields of CH4 and CO2 slightly decreased. The
research conducted by Ref. [59], glycerol was gasified in a fixed value of the LHV, the carbon conversion efficiency, and the gas
bed reactor in the presence of a Ni/Al2O3 catalyst. From their yield were improved due to the increase in the H2, CO, and CH4
results, it was determined that by varying the catalyst loading concentrations. The use of a catalyst also enhanced the tar
from 0 to 0.8 wt%, the composition of H2 gas increased steadily reforming reaction, which decreased the tar yield. In the research
from 55.4 mol% to 68.3 mol%. In another study conducted by Ref. conducted by Ref. [27], a new type of tri-metallic catalyst (nano-
[62], calcined dolomite was used as the catalyst, and the catalyst to NiLaFe/γ-Al2O3) was developed. The effects of this new type of
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1343

catalyst were compared with calcined dolomite and not using any Plus simulation model of the gasification system in bubbling
catalyst. As shown in Table 4, the lowest gas yield of 1.21 m3/kg fluidized bed using air as the fluidization medium and several
and the highest tar yield of 37.8 g/Nm3 were achieved in the different biomasses (olive kernels, corn cobs, and sunflower,
absence of a catalyst. However, when a catalyst was used, the gas rapeseed, and corn stalks) as the feedstock. According to the stu-
yield started to increase and the tar yield decreased. The highest dies conducted by a few different researchers, several assumptions
gas yield of 2.11 m3/kg and the lowest tar yield of 0.28 g/Nm3 were were considered in the modeling the gasification process, as
achieved using the nano-NiLaFe/γ-Al2O3 catalyst. Comparing the follows:
studies with and without catalysts, the concentration of hydrogen
was improved and was able to achieve a value of 53.6 vol% using (i) The processes were considered to be steady state and iso-
the newly developed catalyst. Also with the utilization of a cata- thermal processes [39].
lyst, the CO2 concentration markedly decreased, and the lowest (ii) Instantaneous biomass devolatilization with the volatile pro-
concentration of CO2 was recorded in the process with the newly ducts mainly consisted of H2, CO, CO2, CH4, and H2O [39].
developed catalyst. In the presence of a catalyst, a decrease in the (iii) Within the emulsion phase, all of the of the gases were uni-
concentration of the CO2 was achieved because a majority of the formly distributed [39].
CH4 and C2 hydrocarbons were converted. Table 4, as shown (iv) Particles were spherical in nature with a uniform size and that
below, lists several investigations that utilized different types of the average diameter remained constant during the gasification.
bed materials. This assumption was based on the shrinking core model [39].
From these findings, it can be concluded that, the presence of (v) Char only consisted of carbon and ash [39].
calcium oxide and catalyst in the gasification process can decrease (vi) Sulfur and nitrogen reactions were not considered [12].
the tar yield and increase the total gas yield, H2 gas composition, (vii) In the gas cleaning units, suspended solids were separated
and carbon conversion efficiency. The absorption of CO2 by cal- from the synthesis gas (syngas). Therefore, no solids were
cium oxide is strongly dependent on the partial pressure of CO2 in assumed to be present in any of the downstream units. This
the product stream at the specified gasification temperature. was a simplification of the model because only two fluid
When the equilibrium temperature corresponding to the CO2 phases, liquid and vapor, were considered [44].
partial pressure is higher than the gasification temperature, CO2 is (viii) All of the considered solid species, such as the raw materials,
absorbed and the sorbent gets converted to CaCO3. But when the ashes and char, were treated as Hypo -Components in Aspen
equilibrium temperature corresponding to the CO2 partial pressure Hysys and were defined using their ultimate analysis and
is lower than the gasification temperature, CaCO3 desorbs to their heats of formation [44].
produce the original CaO. The water–gas shift reaction (R7) being
driven in favor of hydrogen in presence of CaO, thus increased the Generally, biomass gasification undergoes the following steps
hydrogen yield. Calcium oxide played a dual role of a sorbent and a in a fluidized bed: (1) the biomass particle decomposes quickly to
catalyst, as the tars and the hydrocarbons are reformed in pre- form char, tar and gaseous products; (2) reactions between the
sence of CaO, thereby producing additional hydrogen [36]. A cat- gaseous products; and (3) tar cracking and char gasification [30].
alyst can increase the gas yields due to the secondary cracking of Table 6 indicates the reactions involved during the gasification
tar in vapor and of hydrocarbons such as CH4 and CnHm in gaseous process [14].
products to generate valuable gases [27]. Besides tar cracking, the Normally, four reactors are needed in a typical production of
tar reforming reactions also increased with presence of catalyst synthesis gas. In the research conducted by Ref. [60], five reactors
which resulted in increased H2 content. were developed using the simulation model in Hysys; this set-up

Table 6
List of reactions in the gasification process [14].
4. Simulation of the gasification process
Reaction Heat of Reaction name Reaction
Experiments, especially on larger scales, are often expensive reactiona number
and complicated; modeling or simulation can save time and
money and it can support the preparation and optimization of Heterogeneous reactions:
C þ0.5O2 ¼CO  111 MJ/ Char partial R1
experiments to be undertaken in real systems [19]. There has been kmol combustion
some research that has been conducted on developing simulations C þCO222CO þ 172 MJ/ Boudouard R2
for the gasification systems. Table 5, as seen below, lists several kmol
studies that have used different types of simulation programs. For C þH2O2COþ H2 þ 131 MJ/ Water–gas R3
kmol
example, in the research conducted by Ref. [42], the simulation of
C þ2H22CH4  75 MJ/kmol Methanation R4
a fluidized bed gasifier using silica sand as the bed material was
developed using Aspen Plus and rice husk was used as the feed-
Homogenous reactions:
stock. In another study, Damartzis et al. [12] developed an Aspen
COþ 0.5O2 ¼ CO2  283 MJ/ CO partial R5
kmol combustion
Table 5 H2 þ 0.5O2 ¼ H2O  242 MJ/ H2 partial R6
Different types of the simulations used in modeling the gasification process. kmol combustion
CO þ H2O2CO2 þ H2  41 MJ/kmol CO shift R7
Types of gasification system Simulation software Reference CH4 þ H2O2COþ 3H2 þ 206 MJ/ Steam-methane R8
kmol reforming
Fluidized bed gasifier Aspen Plus [42]
Fluidized bed gasifier Aspen Plus [12]
Fluidized bed gasifier Aspen Plus [39] Hydrogen sulfide (H2S) and ammonia (NH3) formation reactions:
Pressurized entrained flow gasifier Aspen Hysys and Aspen Plus [44] H2 þ S¼ H2S nrb H2S formation R9
Membrane gasifier Aspen Hysys [16] 0.5N2 þ1.5H22NH3 nr NH3 formation R10
GTI proprietary gasifier model Aspen Hysys [17]
a
Fluidized bed gasifier Aspen Plus [14] Negative sign indicates an exothermic reaction and positive sign indicates an
Fluidized bed gasifier Aspen Plus [40] endothermic reaction.
b
nr ¼ not reported.
1344 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

was chosen because the conversion and equilibrium reactions Table 7


could not be placed in the same reaction chain and therefore could Types of the model used in the simulation programs.
not be placed in the same reactor. The combustor reactor was
Simulation program Types of model used References
separated into a conversion reactor and an equilibrium reactor
[60]. In the Aspen Plus simulation model developed by Ref. [39], Aspen Hysys Equilibrium model [6]
they considered the different stages to show the overall gasifica- Aspen Plus Equilibrium model [14]
tion process, which were (i) decomposition of the feed, (ii) volatile Aspen Plus Equilibrium model [47]
Aspen Plus Kinetic model [12]
reactions, (iii) char gasification and (iv) gas–solid separation. The Aspen Hysys Equilibrium model [4]
Aspen Plus yield reactor, RYIELD, was used to simulate the Aspen Plus Kinetic model [7]
decomposition of the biomass feed into carbon, hydrogen, oxygen,
sulfur, nitrogen, and ash according to the biomass ultimate ana-
lysis. For the volatile reaction stage, the Aspen Plus Gibbs reactor, analysis data [46]. According to Ref. [46], the stoichiometric model
RGIBBS, was used with the assumption that volatile reactions was based on selecting those species that are present in the largest
followed the Gibbs equilibrium. Meanwhile, the Aspen Plus CSTR amounts, i.e., those with the lowest value of the free energy of
reactor, RCSTR, was used to model the char gasification by using formation. Meanwhile, the non-stoichiometric equilibrium model
reaction kinetics [44]. The modeling of the gaseous behavior in was based on minimizing Gibbs the free energy in the system wit-
this gasification system was achieved by using the Peng–Robinson hout specifying the possible reactions taking place.
equation of state [12,40]. Meanwhile, Ref. [16] developed an Aspen Table 7 lists several studies using the different model types. Bas-
Hysys simulation model to study the gasification of three different syouni et al. [6] modeled the date palm waste air–steam gasification in
biomasses (bagasse, switchgrass, and a nutshell mix). The scaling a downdraft gasification unit using a steady state equilibrium model in
factor of 1 was used to determine the hydrogen production from Aspen Hysys. In this model, tars were considered to be non-
an increased feed rate (i.e., doubling the feed rate will double the equilibrium products in order to simplify the hydrodynamics. The
amount of hydrogen produced) [17]. In the Hysys simulation combustion of char and the volatiles in the combustion zone were
model developed by Doong and Lau [16], switchgrass was used as modeled with equilibrium and Gibbs reactor models, respectively. The
the feedstock and the gasifier was simulated as a Gibbs reactor by gasification zone was modeled with a Gibbs and an equilibrium
minimizing the overall free energy of the outlet stream. The reactor. In Aspen Hysys, all of the gasification reactions were modeled
switchgrass ultimate analysis result was used in the simulation as equilibrium reactions except for the oxidation reactions of carbon,
calculation, which involved carbon, hydrogen and oxygen and the which were modeled as conversion reactions. Doherty et al. [14]
other components were ignored [16]. developed a process model for the biomass gasification in a circulating
fluidized bed gasifier using Aspen Plus. The model was based on the
4.1. Modeling of the biomass gasification minimization of Gibbs free energy. RGIBBS reactors allowed restricted
equilibrium specifications for systems that did not reach complete
The modeling of the biomass gasification is important in order equilibrium. Ramzan et al. [47] modeled a kinetic free equilibrium
to study the influence of process parameters, such as the biomass model for the gasification process using Aspen Plus. The partial oxi-
particle mean diameter, the air flow velocity, the gasifier geo- dation and gasification reactions were modeled in an Aspen Plus
metry, the composition and inlet temperature of the gasifying Gibbs free energy reactor. The Gibbs reactor calculated the three phase
agent and the biomass type, on the process propagation velocity equilibrium and syngas composition by the minimization of Gibbs free
(flame front velocity) and its efficiency [55]. Based on the different energy. Tars and other heavy products were assumed to be non-
mechanisms, biomass gasification models can be divided mainly equilibrium products in order to reduce the complexity of the
into two types: a kinetic model and an equilibrium model [30]. hydrodynamics. Damartzis et al. [12] developed a process model for
Some models have utilized the process simulator Aspen Plus and biomass air gasification in a bubbling fluidized bed using the Aspen
have combined the thermodynamic and kinetic rate models [46]. Plus simulator. The model was based on mass and energy balances
Kinetic models are used to predict the progress and the product and the reaction kinetics. In this model, instantaneous devolatilization
compositions at different positions along a reactor. Unlike kinetic was assumed, which neglected the need for the pyrolysis kinetics. The
models, the equilibrium models are used to predict the maximum tar produced from the pyrolysis process was assumed to be a mixture
achievable yield of a desired product from a reacting system. They of three aromatic components, namely benzene, toluene and naph-
can also be utilized as a useful design aid in the determination of thalene, for the facilitation of the development of the Aspen Plus
the possible behavior of a complex reacting system that is difficult simulation model. The tar components were quantitatively deter-
or unsafe to reproduce experimentally or in commercial opera- mined by assuming a 20% w/w conversion of the original biomass to
tions [28,29]. The equilibrium model is known as a zero dimen- tar and a composition of 60%, 20% and 20% for benzene, toluene and
sional model, i.e., a space independent model, and is helpful in naphthalene, respectively. The modeling of the formation of the
identifying the maximum possible conversion of biomass and the volatile gases was performed using an A RGIBBS reactor block in the
theoretical efficiency [1]. However, the equilibrium model cannot Aspen Plus by considering the fact that they followed chemical equi-
be used for reactor analysis and design because it can only predict librium. Alshammari and Hellgardt [4] modeled the gasification of
the end-reaction product distribution and cannot describe the hexadecane using the Peng Robinson equation of state and the direct
instantaneous product distribution along with the geometric minimization of Gibbs free energy in the Aspen Hysys software. This
dimensions [30]. Equilibrium modeling consists of two approa- Gibbs method determined the equilibrium number of moles of the
ches: stoichiometric and non-stoichiometric [1]. The stoichio- reaction components at which the Gibbs free energy of a system was
metric model requires a clearly defined reaction mechanism that minimized. Using this method, Aspen Hysys provided two utilities for
incorporates all of the chemical reactions and species involved modeling the reactions at equilibrium, which are the Gibbs reactor
[46]. This makes this method unsuitable for complex problems and the Equilibrium reactor. In contrast with the Gibbs reactor, which
where the chemical formulas of the feed or the reaction equations required that the reaction components were specified, the Equilibrium
are not well known [1]. In contrast, non-stoichiometric models reactor required the correct stoichiometric data to be specified. These
require no particular reaction mechanisms or species in the Equilibrium and Gibbs reactors were referred to as stoichiometric and
numerical simulation. To specify the feed, the only input needed is non-stoichiometric methods, respectively. Begum et al. [7] developed
its elemental composition that can be obtained from the ultimate a process model for solid waste (wood) gasification in a fluidized bed
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1345

using the Aspen Plus simulator. In this study, both reaction kinetic using waste from the pulp and paper industry. From this research,
parameters and bed hydrodynamic conditions were considered in they determined that 6700 kg steam/ton pulp was produced with
modeling the process. The gasification of char was modeled in a an energy consumption of 5.5 GJ/ton pulp. Chattanathan et al. [11]
continuous stirred tank reactor (CSTR) in Aspen Plus. The reaction used a response surface methodology as a statistical tool for the
kinetics were included in the simulation using FORTRAN and Excel optimization of the conversion of CH4. The maximum CH4 con-
code in the CALCULATOR block. The gasifier was divided into bed and version and the minimum coke formation were achieved at a
freeboard regions using hydrodynamic parameters, where each region temperature of 800 °C, a CO2:CH4 ratio of 1:1 and a CH4:steam
was simulated by one RCSTR reactor. Each RCSTR was then divided ratio of 1:1. Yusup et al. [66] used The Expert Design-8 software in
into a series of CSTR reactors with the same volume. the investigation of the optimization of the production of hydro-
gen with in-situ catalytic adsorption (ICA) steam gasification in a
pilot-scale fluidized bed gasifier. The optimization study was
5. Optimization of the gasification process conducted based on the Response Surface Methodology (RSM)
using the Central Composite Rotatable Design (CCRD) approach.
Optimization can be employed to find the optimal conditions Using this approach, the optimum values (H2 composition and
and/or the geometric parameters that give the extremes of selec- yield of 82.11% and 80.39 g/kg of biomass, respectively) were
ted objective functions in the frame of previously defined con- determined to be a temperature 675 °C, a steam to biomass mass
straints. The objective of an optimization problem can also be the ratio of 2.0, an adsorbent to biomass mass ratio of 1.0, a superficial
maximum product yield or purity, maximum profit, minimum velocity of 0.21 m/s which was equivalent to 4 times the minimum
production time, the best process configuration that yields the fluidization velocity, and a biomass particle size of 1.0–2.0 mm.
maximum operating efficiency, the best combination of raw
materials that generate maximum profit, or the lowest carbon
emission while fulfilling the demand [2]. 6. Economic evaluation of the H2 gas production
The literature that focused on the optimization of biomass
gasification for the production of syngas was quite limited. Table 8, An economic evaluation of the costs and efficiencies is impor-
as shown below, lists some of the studies for the optimization of tant for the implementation of the different scales and technolo-
the gasification. Buragohain et al. [8] reported the optimizations of gies of the various gasification processes [65]. The Gas Technology
the biomass gasification process for the decentralized power Institute [17], who developed the Aspen Hysys simulation model
generation and Fischer–Tropsch (FT) synthesis (a process for the for the production of hydrogen, used the following rules in the
manufacture of synthetic gasoline and diesel) using FACTSAGE scaling of costs in order to be able to adjust to desired size: a
software FACTSAGE. They determined that the optimum range of power law scaling of 0.7 for the handling of solids, the gasification,
the operating conditions for the gasifier for the FT synthesis were: the gas cleaning, the shifting, and the purification systems and
Air Ratio (AR) ¼0.2–0.4, temperature (T) ¼800–1000 °C, and the 0.8 for the power, steam, the balance of plant, the general facilities,
utilization of air as the gasification medium. The optimum ranges the and overall combined cycle system scaling. Drying and steam
of the operating conditions for the decentralized power generation turbine components were scaled at a 0.6 power law. Table 9 shows
were as follows: AR ¼0.3–0.4, temperature¼ 700–800 °C, and the the hydrogen production rates and the resulting economic eva-
utilization of air as the gasification medium. Inayat et al. [23] luations from their analysis. This H2 production cost includes
performed an analysis of the heat integration of the gasification feeding, gasifying, cleaning, reforming, shifting, and purifying
process for the production of hydrogen from EFB; they also con- using Pressure Swing Adsorption (PSA) to produce hydrogen gas
ducted an optimization study for the minimization of the cost of with purity higher than 99.9% [17].
the production of hydrogen using a MATLAB optimization solver Table 10 lists the hydrogen production costs of the previous
named fmincon. The minimum production cost of RM of 5.77/kg studies on the biomass gasification. In the study conducted by Ref.
H2 was obtained at 1150 K, a steam/biomass ratio of 4 and a sor- [38], with regards to the capital cost and operating parameter data,
bent/biomass ratio of 0.87. It was also determined that the the H2 gas production from the empty fruit bunch using a fluidized
hydrogen yield and purity at the optimal condition were bed gasifier was calculated to be 2.11 USD/kg. The capital cost in
0.0179 kg/h and 79.91 mol%, respectively. Lee et al. [26] used the this research included the furnace, fluidized bed gasifier, and the
LINGOv10 software for the optimization of the gasification process construction expenditure. Lv et al. [33] conducted a biomass

Table 8
Optimization studies of the gasification.

Optimization software/method Optimum condition Reference

FACTSAGE AR: 0.2–0.4, T: 800–1000 °C (for FT synthesis) [8]


AR: 0.3–0.4, T: 700–800 °C (for decentralized power generation)
MATLAB Minimum production cost: [23]
RM 5.77/kg H2
T: 1150 K
SB: 4
Sorbent/biomass ratio: 0.87
LINGOv10 6700 kg steam/ton pulp produced with energy consumption of 5.5 GJ/ton pulp. [26]
Response surface methodology Maximum CH4 conversion and minimum coke formation; [11]
T: 800 °C
CO2:CH4 ratio: 1:1
CH4:steam ratio: 1:1
Response surface methodology (the expert design-8 software) H2 composition and yield of 82.11% and 80.39 g/kg of biomass, respectively T: 675 °C [66]
SB: 2.0
Adsorbent to biomass mass ratio: 1.0 superficial velocity: 0.21 m/s

AR: Air ratio; T: temperature; SB: steam to biomass ratio.


1346 A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347

Table 9
Economic results for the gasification of a biomass feedstock [17].

Feedstock Gasifier feed rate Hydrogen Produced Feedstock cost Capital cost H2cost (15% IRR)

Dry tonnes/day Tonnes/day Nm3/day US $/GJ US$ million US$/GJ

Bagasse 400 13.2 347,000 1.50 37.0 10.23


800 62.5 695,000 1.50 61.1 8.74
1600 125 1,390,000 1.50 100.9 7.67
Switchgrass 440 37.0 412,000 1.50 36.5 8.76
880 74.0 824,000 1.50 60.6 7.54
1760 148 1,648,000 1.50 100.9 6.67
Nutshell mix 438 38.7 488,000 1.50 36.3 8.26

Table 10 production costs of hydrogen from different processes. The


Hydrogen production costs in previous studies. hydrogen production cost from gasification process is dependent
on the gasifier design and the process parameters that were uti-
Process H2 production References
cost
lized. The biomass gasification process for hydrogen production
appears to be among the most economic process after supercritical
Empty fruit bunch gasification in fluidized 2.11 USD/kg [38] water partial oxidation. Thus, the hydrogen production from bio-
bed mass gasification has a high potential to be further developed as
Biomass gasification in a downdraft gasifi- 1.69 USD/kg [33]
an alternative to the current hydrogen production process.
cation and a downstream CO-shift reac-
tion in a fixed bed
Biomass steam gasification in fluidized bed 1.91 USD/kg [23]
Supercritical water partial oxidation 0.35 USD/kg [21] Acknowledgments
Electrolysed hydrogen 10 USD/kg [24]

The financial support from the Exploratory Research Grant


gasification in a downdraft gasifier with a downstream CO-shift Scheme, Acc no. 9010-00033 from the MOHE (Ministry of Higher
reaction in a fixed bed. For this system, the calculated H2 product Education) of Malaysia through the School of Bioprocess Engi-
cost was 1.69 USD/kg. Inayat et al. [23] designed a heat integrated neering, Universiti Malaysia Perlis (UniMAP) with the assistance of
flowsheet for the production of hydrogen from oil palm empty the Research Management Centre (RMC) was greatly appreciated.
fruit bunch (EFB) via a steam gasification in a fluidized bed with
in-situ CO2 capture. The minimum production cost of 1.91 USD/kg
was obtained at a temperature of 1150 K, a steam/biomass ratio of References
4 and a sorbent/biomass ratio of 0.87. Hong and Spritzer [21]
developed a Supercritical Water Partial Oxidation (SWPO), gasifi- [1] Acharya B, Dutta A, Basu P. An investigation into steam gasification of biomass
for hydrogen enriched gas production in presence of CaO. Int J Hydrog Energy
cation process involving oxidative reactions in a supercritical 2010;35(4):1582–9.
water environment – akin to high pressure steam – in the pre- [2] Ahmed TY, Ahmad MM, Yusup S, Inayat A, Khan Z. Mathematical and com-
sence of substoichiometric quantities of the oxidant. They esti- putational approaches for design of biomass gasification for hydrogen pro-
duction: a review. Renew Sustain Energy Rev 2012;16(4):2304–15.
mated the cost of the hydrogen produced by this method to be
[3] Alauddin ZABZ, Lahijani P, Mohammadi M, Mohamed AR. Gasification of lig-
approximately 0.35 USD/kg. In another study, according to Ref. nocellulosic biomass in fluidized beds for renewable energy development: a
[24], the cost for the production of electrolysed hydrogen, review. Renew Sustain Energy Rev 2010;14(9):2852–62.
[4] Alshammari YM, Hellgardt K. Thermodynamic analysis of hydrogen produc-
including the capital and operational costs, was 10 USD/kg.
tion via hydrothermal gasification of hexadecane. Int J Hydrog Energy 2012;37
(7):5656–64.
[5] Asadullah M, Ito S, Kunimori K, Yamada M, Tomishige K. Biomass gasification
to hydrogen and syngas at low temperature: novel catalytic system using
7. Conclusions and outlook fluidized-bed reactor. J Catal 2002;208(2):255–9.
[6] Bassyouni M, ul Hasan SW, Abdel-Aziz MH, Abdel-hamid SM-S, Naveed S,
The production of syngas from the gasification technology Hussain A, et al. Date palm waste gasification in downdraft gasifier and
simulation using ASPEN HYSYS. Energy Convers Manag 2014;88:693–9.
offers an opportunity to convert a renewable biomass feedstock
[7] Begum S, Rasul M, Akbar D, Cork D. An experimental and numerical investi-
into a clean fuel gas with a higher yield. Modeling and simulation gation of fluidized bed gasification of solid waste. Energies 2013;7(1):43–61.
of the biomass gasification are the important preliminary studies [8] Buragohain B, Mahanta P, Moholkar VS. Thermodynamic optimization of
biomass gasification for decentralized power generation and Fischer–Tropsch
to predict the effect of process parameters in biomass gasification.
synthesis. Energy 2010;35(6):2557–79.
Aspen Plus and Aspen Hysys are among the tools that can be used [9] Chaiprasert P, Vitidsant T. Effects of promoters on biomass gasification using
to model the biomass gasification based on the application of nickel/dolomite catalyst. Korean J Chem Eng 2010;26(6):1545–9.
[10] Chang ACC, Chang H-F, Lin F-J, Lin K-H, Chen C-H. Biomass gasification for
kinetic model or equilibrium model. From the previous simulation
hydrogen production. Int J Hydrog Energy 2011;36(21):14252–60.
and experimental data, it can be concluded that, reaction tem- [11] Chattanathan SA, Adhikari S, Taylor S. Conversion of carbon dioxide and
perature, ER value and the presence of catalyst are the most methane in biomass synthesis gas for liquid fuels production. Int J Hydrog
Energy 2012;37(23):18031–9.
important parameter in gasification process that significantly
[12] Damartzis TH, Michailos S, Zabaniotou A. Energetic assessment of a combined
determine the gasification performance and the quality of total gas heat and power integrated biomass gasification – internal combustion engine
produced. Meanwhile, the biomass particle size is found to be the system by using Aspen Plus. Fuel Process Technol 2012;95:37–44.
least significant parameter in gasification. [13] Demirbaş A. Biomass resource facilities and biomass conversion processing for
fuels and chemicals. Energy Convers Manag 2001;42(11):1357–78.
Since gasification process requires a significant amount of [14] Doherty W, Reynolds A, Kennedy D. The effect of air preheating in a biomass
energy, optimization study can be employed to seek for the opti- CFB gasifier using ASPEN Plus simulation. Biomass Bioenergy 2009;33
mum conditions of gasification process. Among the useful tools (9):1158–67.
[15] Erakhrumen AA. Biomass gasification: documented information for adoption/
reported for optimization are FACTSAGE, MATLAB, LINGOv10 and adaptation and further improvements toward sustainable utilisation of
Design Expert. This paper also discussed and compared the renewable natural resources. Renew Energy 2012;2012:1–8.
A.A. Ahmad et al. / Renewable and Sustainable Energy Reviews 53 (2016) 1333–1347 1347

[16] Doong SJ, Lau F. Novel membrane gasifier for hydrogen production from bio- [43] Paula A, Peres G, Lunelli BH, Fllho RM. Application of biomass to hydrogen and
mass – Thermodynamic analysis. Gas Technology Institute (Des Plaines). syngas production. Ital Assoc Chem Eng 2013;32:589–94.
[17] Bowen DA, Lau F, Zabransky R, Remick R, Slimane R, Doong S. Techno-eco- [44] Pérez-Fortes M, Bojarski A, Ferrer-Nadal S, Kopanos GM, Nougués JM, Velo E,
nomic analysis of hydrogen production by gasification of biomass. Des Plaines: et al. Enhanced modeling and integrated simulation of gasification and pur-
Gas Technology Institute; 2003. ification gas units targeted to clean power production. Comput Aided Chem
[18] Gil J, Corella J, Aznar MP, Caballero MA. Biomass gasification in atmospheric Eng 2008;25:793–8.
and bubbling fluidized bed: Effect of the type of gasifying agent on the product [45] Pfeifer C, Koppatz S, Hofbauer H. Steam gasification of various feedstocks at a
distribution. Biomass Bioenergy 1999;17(5):389–403. dual fluidised bed gasifier: impacts of operation conditions and bed materials.
[19] Gómez-Barea A, Leckner B. Modeling of biomass gasification in fluidized bed. Biomass Convers Biorefin 2011;1(1):39–53. http://dx.doi.org/10.1007/
Prog Energy Combust Sci 2010;36(4):444–509. s13399-011-0007-1.
[20] Hanaoka T, Yoshida T, Fujimoto S, Kamei K, Harada M, Suzuki Y, et al. [46] Puig-Arnavat M, Bruno JC, Coronas A. Review and analysis of biomass gasifi-
Hydrogen production from woody biomass by steam gasification using a CO2 cation models. Renew Sustain Energy Rev 2010;14(9):2841–51.
sorbent. Biomass Bioenergy 2005;28(1):63–8. [47] Ramzan N, Ashraf A, Naveed S, Malik A. Simulation of hybrid biomass gasifi-
[21] Hong GT, Spritzer MH. Proceedings of the 2002 US DOE hydrogen program cation using Aspen plus: a comparative performance analysis for food,
review NREL/CP-610-32405. Supercrit Water Part Oxid 2002;1:18. municipal solid and poultry waste. Biomass Bioenergy 2011;35(9):3962–9.
[22] Inayat A, Ahmad MM, Mutalib MIA, Yusup S. Biomass steam gasification with [48] Rapagnà S, Latif A. Steam gasification of almond shells in a fluidised bed
in-situ CO2 capture for enriched hydrogen gas production: a reaction kinetics reactor: the influence of temperature and particle size on product yield and
modelling approach. Energies 2010;3:1472–84. distribution. Biomass Bioenergy 1997;12(4):281–8.
[23] Inayat A, Ahmad MM, Mutalib MIA, Yusup S. Heat integration analysis of [49] Rostrup-Nielsen J, Christiansen LJ. Concepts of syngas manufacture. Covent
gasification process for hydrogen production from oil palm empty fruit bunch. Garden: Imperial College Press; 2011.
Chem Eng Trans 2011;25:971–6. [50] Shen L, Gao Y, Xiao J. Simulation of hydrogen production from biomass gasifi-
[24] Iwasaki W. A consideration of the economic efficiency of hydrogen production cation in interconnected fluidized beds. Biomass Bioenergy 2008;32(2):120–7.
from biomass. Int J Hydrog Energy 2003;28(9):939–44. [51] Skoulou V, Swiderski A, Yang W, Zabaniotou A. Process characteristics and pro-
[25] Jansen RA. Second generation biofuels and biomass: essential guide for ducts of olive kernel high temperature steam gasification (HTSG). Bioresour
investors, scientists and decision makers. Weinheim: John Wiley & Sons; 2012. Technol 2009;100(8):2444–51. http://dx.doi.org/10.1016/j.biortech.2008.11.021.
[26] Lee SH, Ng RTL, Ng DKS, Foo DCY, Chew IML. Optimization of a Gasification- [52] Skoulou V, Zabaniotou A, Stavropoulos G, Sakelaropoulos G. Syngas produc-
Based Integrated Biorefinery for Pulp and Paper Industry. In: Proceedings from tion from olive tree cuttings and olive kernels in a downdraft fixed-bed
the 6th international conference on process systems engineering (PSE ASIA). gasifier. Int J Hydrog Energy 2008;33(4):1185–94.
Faculty of Chemical Engineering, Uni (Kuala Lumpur); 2013. [53] Srirangan K, Akawi L, Moo-Young M, Chou CP. Towards sustainable production of
[27] Li J, Yin Y, Zhang X, Liu J, Yan R. Hydrogen-rich gas production by steam clean energy carriers from biomass resources. Appl Energy 2012;100:172–86.
gasification of palm oil wastes over supported tri-metallic catalyst. Int J [54] Tilay A, Azargohar R, Gerspacher R, Dalai A, Kozinski J. Gasification of canola
Hydrog Energy 2009;34(22):9108–15. meal and factors affecting gasification process. BioEnergy Res 2014;7:1131–43.
[28] Li S, Xu S, Liu S, Yang C, Lu Q. Fast pyrolysis of biomass in free-fall reactor for [55] Tinaut FV, Melgar A, Pérez JF, Horrillo A. Effect of biomass particle size and air
hydrogen-rich gas. Fuel Process Technol 2004;85(8–10):1201–11. superficial velocity on the gasification process in a downdraft fixed bed
[29] Li XT, Grace JR, Lim CJ, Watkinson AP, Chen HP, Kim JR. Biomass gasification in gasifier. An experimental and modelling study. Fuel Process Technol 2008;89
a circulating fluidized bed. Biomass Bioenergy 2004;26(2):171–93. (11):1076–89.
[30] Lü P, Kong X, Wu C, Yuan Z, Ma L, Chang J. Modeling and simulation of bio- [56] U.S. Department of Energy National Energy Technology Laboratory. Current
mass air-steam gasification in a fluidized bed. Front Chem Eng China 2008;2 worldwide synthesis gas production. Retrieved from http://www.netl.doe.
(2):209–13. gov/research/coal/energy-systems/gasification/gasification-plant-databases/
[31] Luo S, Xiao B, Guo X, Hu Z, Liu S, He M. Hydrogen-rich gas from catalytic steam current-world.
gasification of biomass in a fixed bed reactor: influence of particle size on [57] Ciferno JP, Marano JJ. Benchmarking biomass gasification technologies for
gasification performance. Int J Hydrog Energy 2009;34(3):1260–4. fuels, chemicals and hydrogen production. U.S. Department of Energy National
[32] Lv PM, Xiong ZH, Chang J, Wu CZ, Chen Y, Zhu JX. An experimental study on Energy Technology Laboratory (Washington, DC); 2002.
biomass air–steam gasification in a fluidized bed. Bioresour Technol 2004;95 [58] United States Biomass Energy Resource Centre. Carbon dioxide and biomass
(1):95–101. energy. Vermont: United States Biomass Energy Resource Centre; 2007.
[33] Lv P, Wu C, Ma L, Yuan Z. A study on the economic efficiency of hydrogen pro- [59] Valliyappan T, Ferdous D, Bakhshi NN, Dalai AK. Production of hydrogen and
duction from biomass residues in China. Renew Energy 2008;33(8):1874–9. syngas via steam gasification of glycerol in a fixed-bed reactor. Top Catal
[34] Lv P, Yuan Z, Ma L, Wu C, Chen Y, Zhu J. Hydrogen-rich gas production from 2008;49(1–2):59–67.
biomass air and oxygen/steam gasification in a downdraft gasifier. Renew [60] Vladan SI, Isopencu G, Jinescu C, Mares MA. Process simulation to obtain a
Energy 2007;32(13):2173–85. synthesis gas with high concentration of hydrogen. UPB Sci Bull Ser B 2011;73
[35] Lv P, Yuan Z, Wu C, Ma L, Chen Y, Tsubaki N. Bio-syngas production from (1):29–36.
biomass catalytic gasification. Energy Convers Manag 2007;48(4):1132–9. [61] Warnecke R. Gasification of biomass: comparison of fixed bed and fluidized
[36] Mahishi MR, Goswami DY. An experimental study of hydrogen production by bed gasifier. Biomass Bioenergy 2000;18(6):489–97.
gasification of biomass in the presence of a CO2 sorbent. Int J Hydrog Energy [62] Wongsiriamnuay T, Kannang N, Tippayawong N. Effect of operating conditions
2007;32(14):2803–8. on catalytic gasification of bamboo in a fluidized bed. Int J Chem Eng
[37] McKendry P. Energy production from biomass (part 1): overview of biomass. 2013;2013:1–9.
Bioresour Technol 2002;83(1):37–46. [63] Wu C, Wang Z, Wang L, Huang J, Williams PT. Catalytic steam gasification of
[38] Mohammed MAA, Salmiaton A, Wan Azlina WAKG, Mohammad Amran MS, biomass for a sustainable hydrogen future: Influence of catalyst composition.
Fakhru’l-Razi A. Air gasification of empty fruit bunch for hydrogen-rich gas Waste Biomass Valoriz 2013;5(2):175–80.
production in a fluidized-bed reactor. Energy Convers Manag 2011;52 [64] Yan F, Zhang L, Hu Z, Cheng G, Jiang C, Zhang Y, et al. Hydrogen-rich gas
(2):1555–61. production by steam gasification of char derived from cyanobacterial blooms
[39] Nikoo MB, Mahinpey N. Simulation of biomass gasification in fluidized bed (CDCB) in a fixed-bed reactor: influence of particle size and residence time on
reactor using Aspen Plus. Biomass Bioenergy 2008;32(12):1245–54. gas yield and syngas composition. Int J Hydrog Energy 2010;35(19):10212–7.
[40] Nooruddin O. Simulation and optimization of IGCC technique for power [65] Yassin L, Lettieri P, Simons SJR, Germanà A. Techno-economic performance of
generation and hydrogen production by using lignite Thar coal and cotton energy-from-waste fluidized bed combustion and gasification processes in the
stalk (Master's thesis). Finland: Lappeenranta University of Technology, UK context. Chem Eng J 2009;146(3):315–27.
Faculty of Technology; 2011. [66] Yusup S, Khan Z, Ahmad MM, Rashidi NA. Optimization of hydrogen produc-
[41] Ollhoff J. Geothermal, biomass, and hydrogen. Minnesota: Abdo Publishing tion in in-situ catalytic adsorption (ICA) steam gasification based on Response
Company; 2010. Surface Methodology. Biomass Bioenergy 2014;60:98–107.
[42] Panda C. Aspen Plus simulation and experimental studies on biomass gasifi-
cation (Bachelor's degree thesis). Rourkela, India: National Institute of Tech-
nology, Department of Chemical Engineering; 2012.

You might also like