Вы находитесь на странице: 1из 29

A Comprehensive Review of Biomass Gasification Process

A.S. El-Shafaya,b, A.A. Hegazia, F. S. H. El-Emama, M. Okashaa


a
Department of Mechanical Engineering, Mansoura University, Egypt
b
Department of Mechanical Engineering, Prince Sattam Bin Abdulaziz University, KSA
Abstract
Gasification can be considered as one of the most conventional techniques in the conversion of biomass because of
some crucial factors, such as the consideration on space, reduction of the substantial waste volume, the flexibility of the fuel
used as well as the recovery of energy. The overall process of Gasification tends to involve the partial combustion of the
organic part of the fossil fuels, which forms an abundant combustible gas in the presence of carbon monoxide, hydrogen, and
some saturated hydrocarbon gases, or methane. When we consider the process of Gasification, the parameters deemed to
affect its performance, including the catalysts, Gasifying Agents or the Biomass ration and temperatures, as well as the type
of raw materials are reviewed and discussed in this study. The primary purpose of this paper is to address the Gasification
Process, including the Types of Gasifiers System Designs like the Fluidized bed, Downdrafts, and the Updrafts. It’s also
responsible for addressing the production of tar from the process of Gasification. The paper explores and discusses all the
primary, secondary and tertiary types of tar while touching on the Gasification Modeling Methods and Equations used in the
Modelling.

Keywords: Biomass, Sawdust, Gasification, hydrogen production and Renewable Energy

(Edit the Table of Contents)


(Edit the Nomenclature for all paper)

1. Introduction
Gasification is a thermochemical process that occurs at high temperatures usually more than 700C,
responsible for converting carbonaceous materials including fossil fuels, biomass, plastics, and coal into syngas which
is consists of H2, CH4, CO, and CO2. A specific amount of oxygen (air) and steam is used as Gasifying agent and
heat carrier agent. The syngas yield can be burnt directly to produce heat at temperatures that are higher compared to
the combustion products. Also, the syngas, if well cleaned, may be applied to different pathways to yield useful outputs
including that:

1. Production of methanol,
2. Purified syngas used in the gas engine, gas turbines, and fuel cells to generate electrical energy,
3. Methane through Sabatier reaction,
4. Production of dimethyl by methanol dehydration,
5. Production of hydrogen, and
6. Production of fuels like gasoline and diesel [32].

This chapter presents a literature summary which involves of six parts: Fluidized bed, an overview of biomass used
in Gasification, tar reduction, Gasification agent, finally influence of Gasifier operating conditions and Mathematical
modeling. Particular attention is paid to literature regarding in tar formation, and mitigation and enhancing of
operational conditions during biomass Gasification are introduced. Finally, refined research requirements for this
research work based on the literature overview are determined.

2. Fluidized Bed Gasifier


In the Gasification process, the biomass is heated to a high temperature producing a series of chemical and
physical changes, which result in the development of Volatile Products plus solid carbonaceous residues. The amount
of the Volatile Products depends on: the heating rate, temperature height, and type of biomass fuel. It is acceptable
that the Char Gasification Stage is the rate-limiting factor in the Gasification of Harvest Residues because of the quick
Devolatilization Stage. The composition of the syngas produced typically depends on the degree of equilibrium
attained by various gas phase reactions especially water gas shift reaction [32, 33].

In the absence of a catalyst, Char Gasification with reactive gases such as oxygen are produced at high
temperatures. Otherwise, when Char is Gasified in the presence of the steam, the gas produced is composed mainly of
CO, CO2, H2, and CH4. In reactor operating at low temperatures, low heating rates and very high-pressure, secondary

-1-
reactions are very significant because of long residence time. At low pressure, high temperature and high heating rates,
most of the Volatile Products escape from the biomass particle due to the pyrolysis process. The process, on the other
hand, reduces the chance of Solid Char reaction with the gas.

In fluidized bed gasifiers, the latter prevails, but because of the mixing nature of the bed, secondary reactions
in the gas-solid and gas phase takes place. Biomass gasification process is considered to be occurring in four stages:
drying of the feedstock pyrolysis to produce volatile matters and char, gasification of the char with reactive gases such
as O2, H2, H2O and secondary reactions of primary gasses and tars [50]. The advantages of the Fluidized bed reactors
are good gas-solid contact, better temperature control, excellent heat transfer characteristics, and high volumetric
capacity. The temperature can be controlled by both of varying the feed rate or the agent rate. Moreover, low operating
temperature leads to produced more slag and clinker. Fluidized bed reactors have the more extensive adaptability to
handle different types of fuel. High ash or moisture content of the feedstock pointing to no problems to the fluidized
gasifiers such as those ordinarily encountered with moving bed Gasifiers.

Figure 1.0: Gasification In Fluidized Beds.

The tar contents of the syngas obtained from Fluidized bed are less than that in the syngas obtained in the
updraft. These desirable features of the Fluidized bed Gasifiers make it more appropriate for large-scale operation than
downdraft Gasifier [35, 50]. The disadvantages of the Fluidized reactors are a significant pressure drop and corrosion
of the reactor body. Because of Fluidized bed reactors operate at pressures little above atmospheric, so must be
designed to prevent leakage. Other disadvantages of the Fluidized bed Gasifier are higher tar content of the product
gas and the incomplete combustion of carbon atoms [37]. Fluidized bed Gasification has been widely used for coal
Gasification for many years; uniform temperature distribution achieved in the Gasification zone also is an advantage
over the fixed bed. The uniformity of the temperature is obtained using a bed of fine material into which air is
introduced, fluidizing the bed materials and guarantee intimate mixing of the hotbed materials, the hot combustion
gas, and the biomass feed. Fluidized bed Gasifiers can be classified by their configuration and the Gasifying agent
velocity into bubbling Fluidized bed, circulating Fluidized bed and spouted Fluidized bed as present in the following
sections.

-2-
2.1. Fluidization and fluidization velocity
The flow rate of oxidizing agent supplied to the fluidized-bed gasifier is a crucial parameter to maintain
effective fluidization of bed materials. The fluidization condition is usually described using a superficial gas velocity
which is the ratio of volumetric gas flow rate and bed cross-sectional area. The superficial gas velocity, at which the
drag force on the bed materials equals the gravitational force, is defined as the minimum fluidization velocity (Umf)
of the bed materials. At minimum fluidization condition (shown in Fig. 3.3)., the bed materials lift upward and remain
in suspension; bed pressure drop (dTmf) reaches to a maximum and remains constant with further increase in the
superficial gas velocity. Fluidization characteristics, such as Umf and dPmf, depend upon the particle size and
composition of the bed materials [137].

Umf and dPmf are also influenced by segregation and mixing behaviors of bed materials. Segregation is a process
during which a bed material with higher particle density, such as sand, moves downwards in the bed while a material
with lower particle density, such as biomass, floats upwards [138,139]. This, in turn, causes separation of biomass
from sand and results in a localized accumulation of biomass particles as smaller and more significant sized lumps
throughout the bed. These lumps further lead to channel formation, called in bed channelization, that give rise to larger
void space and a shorter path to the gas flow [140]. As a result, the gas quickly escapes through in-bed channels, which
affect bubble formation, and thus turbulence level in the bed resulting in ineffective fluidization. Segregation occurs
due to differences in densities or sizes of the bed materials such as sand and biomass [141, 142]. Genially, when a
packed bed of particles is subjected to a sufficient high upward flow of fluid the weight of the particles is supported
by the drag force exerted by the fluid on the particles and the particles become freely suspended or fluidized. The
behavior of fluidized suspension is similar in many aspects to that of a pure liquid. Mass transfer and heat transfer
rates between particles and submerged objects (e.g., heat exchanger tubes) are significantly enhanced in fluidized beds,
to avoid or reduce carryover of particles from the fluidized bed, keep the gas velocity between minimum fluidization
velocity (Umf) and terminal velocity (Ut). Also, rapid particle mixing allows uniformity in bed. As a result, the
fluidized bed is widely used for conducting gas-solid reactions (coal combustion), gas-solid Catalytic Reactions
(catalytic cracking of petroleum), etc. [142].

Therefore, the quantity of biomass in the mixture plays a crucial role in segregation behavior of bed materials. Also,
a bed consisting of a material, such as particulate matters, that has adhesive or cohesive properties may enhance
segregation tendency and suppress fluidization [143]. Chok et al. [144] indicated improved mixing with a decrease in
the particle size ratio from 30 to 20 of palm shell and sand mixture. The author also reported that segregation and
channelization were predominant at higher particle size ratio and biomass weight fraction (10% and 15%) in the
mixture. Correlations to determine Umf and bed expansion of coal particles were suggested for coal gasification [145].
These methods are derived from pressure drop method, dimensional analysis, drag force method, and terminal velocity
method. At the onset fluidization, drag force by upward moving gas on the whole system of particles must be equal to
the weight of particles of the bed. The range of fluidizing velocity, Umf, in a fluidized bed should be within the
minimum fluidization and terminal velocities of the mean bed particles. In the following subsection the Minimum
fluidization velocity, the Terminal velocity of the particle, Fluidization velocity during the gasification and agent flow
rate to meet the fluidization condition calculation procedures are described as following.

-3-
Fig. 1.1 Fluidization characteristics of sand

2.2. Bubbling Fluidized Bed Gasifier

The bed is termed a bubbling fluidized bed when the granular material (e.g., sand) is lifted in a reactor through
which an upward flow of gas is passing through it at a flow rate where the pressure drop across the particles is enough
to support their weight. In bubbling fluidization, low fluidization velocity just above the minimum agent velocity
passes through the bed in the form of bubbles. Bubbling bed gasifiers consist of a vessel with distributor plate at the
bottom through which the air is introduced. Above the distributor plate the moving bed of fine-grained material into
which the prepared biomass comes in. The setting of the bed temperature to 600-1000C is maintained by controlling
the equivalence ratio. The biomass is pyrolyzed in the hotbed of char form and gaseous compounds, the high
compounds weight being cracked by contact with the hotbed materials, giving a product gas with low tar content. The
bubbling bed gasifiers can be classified according to the number of the bed, such as single fluidized bed and multi
fluidized bed [38].

Fig 1.2: Bubbling Fluidized Bed Gasifier.

2.2.1. Single Fluidized Bed Gasifiers


This system consists of one bed, into which the feedstock and the gasifying agent enter and out of which the
produced syngas and char exit. The advantage of this gasifier is, low operating and initial cost compared to the multi-
bed, less maintenance, and maintenance cost due to uses of one bed only, the produced syngas is ready to direct use.
On the other hand, the disadvantages of single bed as generated syngas heating value is lower than that provided by
the double bed; pyrolysis process occurs in the combustion zone at the bottom of the gasifier which leads to non-
uniform temperature distribution and the separation of the inorganic materials in the feedstock is hopeless.

-4-
Figure 1.3: Single Fluidized Bed Gasifiers

2.2.2. Dual and Multi-Fluidized Bed Gasifiers


These gasifiers consist of more than one bed commonly first bed used to burn some of the char to produce
the energy for the second bed in which the pyrolysis process occurs. The advantage of the dual bed system was reported
by many authors [37]. These include the syngas heating value more significant than that produced by the single bed
due to the combustion of the char occurs in separate reactors and hence the combustion syngas does not dilute the
pyrolysis gases and separation of the inorganic material in the feed is possible. But, the disadvantages are higher
construction cost and more maintenance requirement as compared to the single bed. Keiichi Tomishige et al. [37]
studied the gasification of the biomass by using the Single bed and Dual bed gasifier. It was found that in the Dual bed
gasifier combined with a suitable catalyst; some the tar can be converted to syngas at a lower temperature than that
needed by the traditional methods with high energy efficiency.

-5-
Figure 1.4: Dual Fluidized Bed Gasifiers

2.3 Circulating Fluidized Bed Gasifiers

If the gas velocity in a bubbling bed is too increased, more particles will be entrained in the syngas stream
and leave the gasifier. Finally, the transport velocity for the most of the particle is reached, and the vessel can be
quickly empty of the solid except new particles are fed to the base of the gasifier. If the particles leaving the vessel
return through an outer collection system, then the gasifier now called a circulating fluidized bed. The Circulating
Fluidized Bed has high processing capacity compared to the conventional gasifiers, better gas-solid contact and the
ability to handle coherent solids that otherwise be difficult to fluidize in a bubbling bed. Regardless of these
advantages, the circulating fluidized bed gasifiers are yet less conventional than the bubbling ones because of their
height which restricts their application regarding cost analysis [39].

-6-
Figure 1.5: Circulating Fluidized Bed Gasifiers

2.4 Spouted bed gasifier


A spouted fluidized bed gasifier consists of a bed of a solid particle partially filling the reactor which is
provided with a relatively large control opening at its base. Gas is injected through this aperture. With sufficient flow
of the gas, the particle in the gas can be made to rise in a jet at the center of the bed and develop periodic motion on
the bed as a whole. The bed motion can be assisted by the extra air at the base to produce a spouted bed. Spouted bed
gasifiers have been used to gasifying the coal of various ranks. Abdul Salam [40] has studied the gasification of
biomass in a spouted bed, which has specific possible advantages over the fluid bed configurations. They found that
at a higher agent velocity, the gasification efficiency of circular slit spouted bed was slightly high compared with that
of central jet spouted.

-7-
Figure 1.6: Spouted bed gasifier

3. Overview of Biomass Used in Gasification


In this section, several results from the gasification studies are summarized to present the biomass gasification
process utilization and its performance.

3.1. Crop Residues Gasification (insert the summary of last 3 researches related to this title done in 2016, 2017
and 2018)
Chen et al. [41] perform some experiments involving the gasification of biomass blends in a bench scale
fluidized bed gasifier at atmospheric pressure. Two types of mixtures were prepared, which mixing of pine chips with
low-grade black coal and Sabero coal in the ratio range of 0:100-100:0. The experiments were conducted using
gasification agent as mixtures of air and steam at gasification temperatures of 840C-910C and fluidized gas velocities
of 0.7-1.4m/s. The results are showing that efficiently blending improved the performance of fluidized bed gasification
of the low-grade coal, and the possibility of converting the refuse coal to a low heating value syngas. Their study
indicated that a blend ratio with no less than 20% chip pellets for the low-grade coal and 40% chip pellets for the
refuse coal are the most suitable. The low heating value of the Dry Product Gas increase with increasing blend ratio
from 3.7 to 4.5 MJ/Nm3 for chip pellets /low-grade coal and from 4.0 to 4.7 MJ/Nm3 for chip pellets/refuse coal. The
syngas yield increase with the increase of the blend ratio from 1.80 to 3.20 Nm3/kg (chip pellets /low-grade coal), and
from 0.75 to 1.75 Nm3/kg (chip pellets /refuse coal), respectively. Also, about 50% thermal efficiency was achieved
for the two types of the blend.

3.2. Straw Gasification (insert the summary of last 3 researches related to this title done in 2016, 2017 and 2018)
Sadaka et al. [42] develop a fluidized bed gasifier to gasify wheat straw in an air/steam mixture as an agent.
Their study showed that the performance of the fluidized bed gasifier (gasifier temperature, pressure drop, heating
value and syngas flow rate) was influenced by fluidization velocity, steam (agent) flow rate, and biomass/ steam ratio.
The temperature of the gasification bed reached its maximum at of 891C, indicating that the agglomeration problem
was overcome by injecting more air/steam mixture. The heating value of the syngas yield and the syngas flow rate
reached their maximum of 10.63MJ/Nm3 and 2.47Nm3/min, respectively. The bed pressure drop occurs ranged from
37.2 to 51.6 cmH2O, gives a good fluidization quality. Risen's et al. [43] studied the influence of calcium addition in

-8-
straw gasification. The effect of calcium addition as calcium sugar/molasses solutions to straw significantly influenced
the ash chemistry and the ash sintering tendency but much less the char reactivity. Walawender et al. [44] studied the
straw gasification with steam as an agent in a 0.23m diameter fluidized bed gasifier over a temperature range of 552
to 757C. The fraction of the biomass was converted to syngas ranged from 32% at the temperature of 552C to 73% at
the temperature of 757C. The heating value of the syngas appeared at parabolic variation with temperature with a
maximum value of 12.3MJ/Nm3 obtained at 672C.

3.3. Rice Husk Gasification (insert the summary of last 3 researches related to this title done in 2016, 2017 and
2018)
Yin et al. [45] tested rice husk gasification in circulating fluidized bed for biomass gasification and power
generation system to provide power for a rice mill. The system consists of a circulating fluidized bed gasifier, a gas
cleaner (including an inertial separator, a cyclone, filter, a venture and two water scrubbers). It was found that the
gasifier can be operated stably within the temperature range of 700C to 850C; it's optimal condition. The main
performance indices at capacity, 1500 kg/hr., are gasification efficiency 65%; rice husk consumption, 1.7-1.9kg/Whr;
total efficiency, about 17%. The gasification of rice husk was also studied by Chen and Rei [41] within a temperature
range of 600C to 700C. They used electrical heaters to provide the heat required to gasification in a 0.05m internal
diameter fluidized bed reactor; the bed consisted of alumina sand and agent was super-heated steam. The syngas yield
increased from 0.38 to 0.55m3/kg and the heating value varied from 12.8 to 18.5MJ/m3. Over this temperature range
H2, CH4, CO, and CO2 concentrations in the syngas produced changed from 3.6 to 13.1%, from 14.4 to 13.5%, from
52.2 to 51.1% and from 23.0 to 14.6%, respectively.

3.4. Cotton Gin Trash Gasification (insert the summary of last 3 researches related to this title done in 2016,
2017 and 2018)
Some agricultural waste as biomass samples such as rapeseed, cotton refuse, sunflower shell, pinecone, and
olive refuse was the first pyrolysis in air. Their chars were then gasified in a gas mixture of steam and air. Pyrolysis
of the biomass samples was performed at a heating rate of 20K/min from ambient to 1273K in a dynamic air
atmosphere of 40cm3/min. It was concluded that gasification characteristics of biomass chars were somewhat
dependent on the biomass properties such as ash and fixed carbon contents and the constituents present in the ash.
Singh et al. [46] performed the steam gasification of cottonwood branches in a fluidized bed gasifier and compared
the produced syngas characteristics with those for pure cellulose. The results showed that the syngas heating value,
the energy recovery, carbon conversion and mass yield of the syngas obtained from cottonwood were found to be
lower than those derived from pure cellulose. Groves et al. [47] studied the gasification of cotton gin trash in a fluidized
bed with air over the temperatures range of 649C and 87C in a 0.3m internal diameter gasifier. Its results showed the
syngas heating value and the energy recovery increased from 3.4 to 4.3MJ/Nm3 and from 27 to 53%, respectively.

3.5. Corn Cobs Gasification (insert the summary of last three types of research related to this title done in 2016,
2017 and 2018)
Results for the gasification of corncobs over a temperature range of 500C to 1000C were presented by Epstein et
al. [48]. Produced syngas yield (reported as a mass fraction of the biomass fuel feed) increased from 0.17 to 0.60kg/kg.
The syngas contains large amounts of CO and H2 and the heating value varied from 1.4 to 10.9 MJ/m3.

3.6. Sawdust Gasification (insert the summary of last three types of research related to this title done in 2016,
2017 and 2018)
Wander et al. [49] observed that, the technology of wood gasification could produce a syngas’s qualified of being
combusted in an internal combustion engine, as long as it is appropriately cleaned enough from tar and ashes. In order
to evaluate the performance of the woody biomass gasification process, a small downdraft, fixed bed, stratified and
open top gasifier was constructing. This gasifier, whose capacities were around 12kg/hr., has an internal gas
recirculation, new to this type of gasifier, which can burn part of the syngas produced to raise the gasification reaction
temperature. The syngas yields ranged from 1.1m3/kg at 600C to 1.3m3/kg at 800C, and the heating value was more
than 11.2MJ/Nm3 for all temperatures. Lian and Findley [34] tested the air gasification of oak sawdust and found that
the tar and char yields decreased linearly with temperature from 6% of the dry wood weight at 650C to 0.5% at 800C.
They tested the effect of equivalence ratio on the gasification performance and concluded that the total carbon-to-air
ratio in the dry syngas gave the better correlation with the concentrations of carbon, hydrogen and the higher heating
value.

-9-
3.7. Cellulose Material Gasification (insert the summary of last three types of research related to this title done
in 2016, 2017 and 2018)
Walwender et al. [45] performed alpha cellulose gasification in a bench scale fluidized bed gasifier with steam as
an agent within a temperature range of 600C to 800C. The significant components of the produced syngas were H2,
CO2, CO and CH4 and the volumetric syngas yields were 0.5-1.4m3/kg, the average gas heating value was
11.8MJ/Nm3. The energy recovery and carbon conversion were within the range of 32-90% range. Hoveland et al.
[50] studied the steam gasification of alpha cellulose at high temperatures and confirmed an earlier assumption that
proposed the existence of two temperature regimes for cellulose gasification, one below 567C dominated by volatile
cracking and the other above 567C dominated by the water gas shift reaction.

3.8. Manure Gasification (insert the summary of last three types of research related to this title done in 2016,
2017 and 2018)
Sweeten et al. [51] reported that, feedlot manure has approximately half the heating value of coal, twice the volatile
matter of coal, four times the N content of coal on heat basis. Due to soil pollution during collection, the ash content
is roughly 9-10 times that of low ash (5%) coal. The addition of 5% biomass residues had a little apparent effect on
heating value. Based on heating values and alkaline oxides, partial composting seems favorite to a full composting
cycle. Even though the percentage of alkaline oxides is reduced in the ash, the increased total ash percentage results
in an increase of total alkaline oxides per unit mass of fuel. Raman et al. [36] tested the gasification of feedlot manure
with different fluidization velocities. They found that the fluidization velocities did not have a significant influence on
produced syngas yield, heating value or composition. Walawender and Fan [52] studied the feedlot manure gasification
using air as an agent, found that the generated syngas yield, the energy recovery and the higher heating value increased
by 131%, 244% and 77% when the temperature increased from 627 to 827C. Halligan et al. [53] gasified feedlot
manure in a 0.05m internal diameter fluidized bed gasifier. The agent was a mixture of air and steam, and the bed
consisted of the feed material only. Over the temperature range of 693C to 796C, the syngas yield flow rate increased
from 0.6 to 1.3m3/kg, and the gas heating value increased from 8.7 to 9.8MJ/Nm3. The energy recovery and carbon
conversion also risen from 23 to 49% and from 20 to 50%, respectively.

4. Tar Reduction
Tar is one of the most complex compounds produced from biomass gasification, and its formation is strongly
dependent on the operating conditions such as gasification temperature, a gasifying agent used and equivalence ratio.
Therefore, to avoid various problems related to tar condensation and formation of tar dust, tar reduction is necessary
before the final use of the product syngas. In the following, detailed presentation about tar definition, tar measuring as
well as how to reduce it.

4.1. Tar Definition


Due to its complication and different definitions have been given by various researchers working on biomass
gasification. In Milne’s review report, tar is defined as the organics produced under thermal or partial oxidation
regimes of any organic material and assumed to be hugely aromatic [54]. The European Committee agreed on
unanimity on the definition of tar for Standardization (CEN), and tar is defined as all organic compounds existing in
the gasification product gases with molecular weight more than benzene [55, 56]. During biomass gasification process,
tar is formed in a series of complex reactions and its formation is strongly dependent on the process conditions. The
tar formation layout proposed by Elliott and summarized by Milne [57], which shows the transition of tar as a function
of process temperature from primary products to phenolic compounds to aromatic hydrocarbons. Evans and Milne
[54] reported that the tar could be classified into four product classes which are identified as a result of gas-phase
thermal cracking reactions:

1. Primary products which are characterized by cellulose-derived, hemicellulose-derived and lignin-derived products;
2. Secondary outcomes which are described by phenolics and olefins;
3. Alkyl tertiary products which are mainly methyl derivatives of aromatic compounds; and
4. Condensed tertiary products which are PAH series without a substituent.

The composition, quantity, and properties of tar in product syngas vary depending upon the biomass feedstock,
gasification agent, gasifier type and gasification conditions. Among them, the type of gasifier is one of the fundamental
parameters influence the tar content. The downdraft fixed bed (DFB) was the favored effective method in suppressing

- 10 -
the formation of tar during biomass gasification. For different utilization applications, the maximum allowable tar
content is different. For instance, Boerrigter et al. [58] reported that for Fischer-Tropsch (FT) Diesel synthesis the
naphthalene concentration in product syngas should be less than 2ppm to avoid condensation during the compression
stroke before catalytic conversion. Bui et al. [59] mentioned that the preferable tar and dust loads in syngas’s for
engines must be lower than 10 mg/Nm3.

4.2. Tar Reduction (insert the summary of last 3 researches related to this title done in 2016, 2017 and 2018)
Devi et al. [60] reported that tar removal technologies could be divided into two types: primary methods and
secondary methods. The primary purpose includes (a) the proper selection of the operating conditions; (b) the use of
a suitable bed additive or a catalyst during gasification; and (c) an appropriate gasifier design. Secondary methods
consist of chemical (catalytic cracking) or physical (mechanical separation) treatment [61]. Concerning this two
methods, Corella et al. [62] perform a comparison between them; Corella found no significant difference in their
influence concerning tar reduction. Sutton et al. [63] reported that a suitable combination of different primary and
secondary treatments is likely to improve gasifier performance and produced syngas with minimum tar concentration.
For both methods, tar decomposition mainly occurs due to a chain of complex, multiple and simultaneous reactions
such as steam and dry reforming responses, cracking. Although secondary methods are reported to be very functional
in tar reduction, in some cases, they are not economically viable [62]. Since they are out of the scope of this research,
detailed information is available in the excellent reports by Milne et al. [57] and Neeft et al. [64]. Primary methods
are gaining a lot of attention as they may eliminate or profoundly reduce the need for downstream cleanup. Widely
studies have been conducted by different researchers regarding effects of operating conditions (e.g., steam to fuel ratio
(SFR), temperature, pressure and residence time) and active materials (e.g., bed material, additive) on tar formation
during biomass gasification. A proper selection of operating conditions can strong reduce the amount of tar produced.
Moreover, the optimized operating conditions using catalyst during biomass gasification can also promote char
reaction, reduce the tar yield.

4.2.1. Effect of Temperature on Tar Reduction (insert the summary of last 3 researches related to this title done
in 2016, 2017 and 2018)
The temperature is one of the most critical factors affecting the overall biomass gasification process. Temperature
can influence the amount of tar formed as well as the composition of tar. Kinoshita et al. [65] reported that the total
number of detectable tar species produced from sawdust gasification decreased as well increasing of temperature.
Lower temperatures favored the formation of more aromatic tar species with diversified substituent groups, while
higher temperatures supported the establishment of fewer aromatic tar species without substituent groups. Li et al. [66]
reported that the amount of tar gain from biomass gasification strong decreased from 15 to 0.54 gram/Nm3 as the
average temperature increased from 700C to 820C. Van Paasen and Kiel [67] observed that tar concentration decreased
within a temperature range from 750C to 950C, at the same time the tar compositions shifted from alkyl substituted
poly-aromatic hydrocarbons (PAHs) to non-substituted PAHs. The effect of gasification temperature on tar
concentration is shown in figure (2.1), from which it can be seen that gasification temperature mainly affects the
formation and composition of tar. Kurkela et al. [68] and Simell and Leppälahti [69] studied the effects of operating
conditions on the creation of tar produced from different feedstocks ranging from hard coals to wood gasification in a
pressurized fluidized bed gasifier. They found that, the total tar concentration in the pressurized fluidized bed product
syngas appears to depend mainly on the feedstock and on the gasification temperature, which can be clearly seen in
figure (2.2) which shows the tar content produced from for wood, peat and brown coal gasification in the pressurized
fluidized bed gasifier at different freeboard temperatures.

4.2.2. Effects of Equivalence Ratio and Steam to Fuel Ratio on Tar Formation (insert the summary of last 3
researches related to this title done in 2016, 2017 and 2018)
Similar to temperature, an increase in equivalence ratio (ER) also has a useful effect on reducing tar yield. Narvaez
et al. [71] studied operating conditions on the product syngas produced from sawdust pellets gasification with air in
an atmospheric bubbling fluidized bed gasifier. The effect of ER on the tar yield in the syngas is shown in figure (2.3)
for two H/C ratios in the bed. The tar content produced from sawdust pellets gasification at 800°C decreased with
increasing ER, the tar content was found at about 2-7 gram/Nm3 obtained at an ER value of 0.45. Also, the H/C ratio
is also almost critical, and the tar content decreases with increasing H/C ratios. Lv et al. [72] reported that the lower
heating value (LHV) of the product gas decreased with an ER increase due to intensification oxidization reactions of
product syngases. Steam to fuel ratio (SFR) also influences tar formation due to more(less) tar steam reforming
reactions. Herguido et al. [73] observed that the amount of tar decrease in sharply trending from 8% wt. to minimal

- 11 -
content with an increasing SFR range from 0.5 to 2.5. Aznar et al. [74] reported that with varying the ratio of
((steam/oxygen) / biomass mass) from 0.7 to 1.2 more than 85% reduction in the total tar was recorded.

Figure (2.1) Effect of gasification temperature on tar Figure (2.2) Effect of free-board temperature on the
concentration and dew-point [67] total amount of tar in the pressurized fluidized bed
gasification with different feed-stocks [68]

5. Effect of Agent Type on Gasification


Biomass gasification using fluidized bed has been performed in various gasifying agents such as air, steam,
pure oxygen, carbon dioxide and hydrogen [75]. In the following subsections, a detailed discussion of agent types and
their applications are presented.

5.1. Air Gasification (insert the summary of last three researches related to this title done in 2016, 2017 and
2018)
Gasification process using air as a gasifying agent is most straightforward and widely used method. Excess char
formed by the pyrolysis process within the gasifier is burned with a limited supply of air (usually at ER=0.25). The
syngas has a lower value of the heating value due to the dilution effect of nitrogen from the air. The heating value of
the produced syngas is in the range of 3.5-7.8MJ/Nm3, which makes it suitable for boiler and engine applications.
Due to its simplicity, air gasification technology is being studied by many researchers for various types of biomasses.
Because air is the gasifying agent, the reactor temperature is dependent on the air flow rate and/or biomass feed rate.
Shallow inlet air to the system results
Figurein (2.3)
superficial bed temperature,
Tar concentration whichequivalence
at different produces lower syngas, and higher tar
content. Groves et al. [47] tested the fluidized bed ratioairvalues
partial
at oxidation
800°C [70]of the cotton gin trash within temperatures
range of 922K and 1144K in a 0.3m internal diameter reactor. The syngas yield heating value increased from 3.4 to
4.3MJ/m3. Lian and Findley [34] studied the oak sawdust gasification with air as agent; the results showed that the tar

- 12 -
and char yields decreased as temperature increased by 6%. The researchers tested the effect of ER on the gasifier
performance and concluded that the total carbon/nitrogen ratio in the dry syngas gave the better correlation with the
concentrations of carbon, hydrogen and the higher heating value. Walawender and Fan [50] studied the air gasification
of feedlot manure, were observed that the syngas yield, heating value, and energy recovery increased by 131%, 77%,
and 244%, respectively; when the temperature increased from 900K to 1100K.

Ergudenler [75] tested the effect of ER on the syngas quality and its flow rate in the air as an agent of wheat straw
through a fluidized bed gasifier. The results showed that, at ER of 0.25, the mole fraction of the combustible syngas
achieved their maximum. Cao et al. [91] studied the sawdust gasification in a fluidized bed using air. They combined
two individual zones of pyrolysis, gasification, and combustion of biomass fuel in one gasifier. The primary air stream
and the biomass fuel were interfering to the gasifier from the bottom and the top, respectively. Secondary air was
injected into the upper zone of the reactor to maintain a high temperature. The study indicated that under optimum
operating conditions; syngas produced at a rate of 3.0 Nm3/kg and heating value at about 5.0 MJ/Nm3. The
concentration of carbon monoxide, hydrogen, and methane in the fuel gas produced were 9.25%, 9.27%, and 4.21%,
respectively.

5.2. Steam Gasification (insert the summary of last three researches related to this title done in 2016, 2017 and
2018)
Steam gasification requires an external heat source mainly when used as a gasifying agent, unlike of air gasification.
Using a mixture of steam/air as a gasifying agent is not unique technology; in fact, this mixture has been studied by
many researchers. The oxygen present in the air helps to provide the required energy due to the exothermic nature of
biomass fuel burning. Also, the high temperature assists the devolatilization process of biomass to produce various
gases. Steam reacts with carbon monoxide to produce hydrogen and carbon dioxide, this reaction namely water-gas-
shift reaction. By comparison to air gasification, steam gasification generates a higher energy content producer syngas.

Boateng et al. [77] studied the effects of reactor temperature and SFR on product syngas composition and heating
value. The syngas yield contains a high concentration of hydrogen; have heating value ranged from 11.1MJ/Nm3 at
700C to 12.1MJ/Nm3 at 800C. Also, energy recovery varied from 35%-59% within the same temperature range.
Walwender et al. [50] studied corn grain-dust gasification in a 0.05m internal diameter fluidized bed gasifier by steam
as agent and a mixture of sand/limestone as the bed material. The produced syngas increased from 0.13m3/kg at 867K
to 0.73m3/kg at 1033K. The syngas heating value rose from 9.4MJ/Nm3 to 11.5MJ/Nm3 in the examined temperature
range. The main components of the syngas were H2, CO2, CO, and CH4 and the syngas flow rate yield was 0.5-1.4
m3/kg, heating value was 11.8MJ/m3and the carbon conversion within range of 32-90%.

According to Slapak et al. [78], steam gasification is one of the possibilities for recycling waste in fluidized bed reactor.
The primary product is syngas, applicable for energy recovery. The produced syngas has a heat value of 8.6MJ/Nm3.
Mermoud et al. [79] perform experimental gasification of charcoal using steam of beech charcoal spheres of different
diameters (10-30mm) at different temperatures (830C-1030C). The results showed that, a prolonged reaction at 830C.
A difference in gasification rate as high 6.5 to 1 was observed between temperatures at 1030C and 830C. Experiments
carried out with mixtures of H2O/N2 at 10%, 20%, and 40% mole of steam proven that, oxidant partial pressure
influences gasification. A gasification rate of 1.9 was gain for H2O partial pressure varying from 0.4 to 0.1 bar.

Corella et al. [73] observed on steam gasification of four different biomasses (wood chips, thistle, sawdust, and straw)
in a 0.15m internal diameter fluidized bed gasifier. They determined the syngas produced, char and tar yield at
temperatures ranged from 650C-780C for each type of biomasses. Straw and sawdust showed higher syngas and lower
tar yields compared to wood chips and thistle. Wei et al. [80] tested the legume straw and sawdust pellets as biomass
in downdraft gasifier by varying the operating conditions during experiments; the examined factors affecting the
hydrogen concentrations are detected.

5.3. Oxygen Gasification (insert the summary of last 3 researches related to this title done in 2016, 2017 and
2018)
If the amount of nitrogen supplied to the gasification process is restricted, the product syngas will not contain
nitrogen thus, has average energy content (about 12-21MJ/Nm3). Such the syngas can be distributed in pipeline
network systems at lower cost, therefore, be appropriately used for process heat or possibly as synthesis gas fuels. In

- 13 -
this case, oxygen plant or nearer sources of oxygen are required, this maybe elevates the initial cost needed for the
plant installation.

Bailie [81] recoding the relationship between the heating value of a syngas and oxygen concentrations in the agent. At
oxygen concentration of 20% and 100%, the heating values of the syngas yield were 6 and 11.2 MJ/Nm3, respectively.
At the level of oxygen in the agent increased from 20% to 100%, the methane, hydrogen, and carbon monoxide mole
fraction increased from 4 to 6%, from 13 to 19% and from 25 to 55%, respectively. Tillman [82] studied the municipal
solid waste gasification using oxygen. The biomass fuel (shredded and magnetically sorted) was fed into the top of
the gasifier, and the oxygen was supplied at the bottom. Char was burned by the assist of oxygen at the bottom of the
gasifier to produce enough energy at temperatures in the range of 1593C-1704C and to yield a molten slag from all
noncombustible materials. The maximum concentrations of the produced syngas for CO, H2, CO2, and CH4 detected
were 44%, 31%, 13% and 4%, respectively. The maximum low heating value was 10.6MJ/Nm3.

Watkinson et al. [83] gasified coal to study the effect of an oxygen/steam mixture on carbon conversion and syngas
low heating value during the gasification in a spouted bed. They record that, by increasing the ratio of oxygen/biomass
ratio from 0.5 to 1.1kg/kg, the low heating value of the syngas yield increased sharply from 5MJ/kg to 16MJ/kg.
Biogas dry reforming is a catalytic process that operates at 500C-800C to produce gas that simultaneously consumes
two main greenhouse gases, CH4 and CO2, therefore supporting environmental preservation efforts [84]. The ratio of
the generated H2/CO is unity in stoichiometric reaction, which is suitable for further use in hydro-formulation and
carboxylation process to synthesize liquid fuels [85]. Therefore, dry reforming has become a better option for
renewable biogas energy, although, several problems might prevent its further industrial application. The primary
attention is energy consumption [86]. Because dry biogas reforming is a highly endothermic reaction, for the dynamic
thermal significance, a higher temperature is necessary to gain significant conversions. Another attention is the active
catalyst suppression, mainly caused by the carbon sedimentation and active metal sintering [87], mainly when using
base catalysts, which will also increase the operating cost of biogas reforming process. Finally, biogas is consisting of
some gas species, and the effects of components besides CH4 and CO2, such as O2 have not been enough study on
biogas reforming. Studies have observed that two primary reactions can be caused by carbon formation which
including of CH4 decomposition and CO disproportionation [88]. On another hand, a reaction that often occurs during
dry biogas reforming is reverse water gas shift (RWGS) reaction which consumes more H2 and produces more CO.
Therefore, the decreasing in the H2/CO ratio occurs [89]. In the presence of oxygen, partial CH4 oxidation and total
CH4 oxidation, carbon oxidation, CO oxidation and H2 oxidation may occur, where carbon gasification can help to
remove carbon sedimentation.

The process of combining of dry and partial oxidative reforming of methane, also known as auto-thermal reforming
of methane, have been recorded to be useful in recover energy consumption of dry reforming and adjusting the H2/CO
ratio in syngas yield. The use of pure oxygen and steam required a significant amount of input energy and expensive
initial investments [90]. Butterman and Castaldi studied the production of more active chars formed by the heat
treatment of woody biomass as fuel with CO2, compared to those developed using steam [90]. Seiler et al. suggested
that it’s possible to modify the economics of the biomass to liquid process through the sale of N2 which formed as a
by-product from the production of oxygen enriched air [90]. The gasification process with oxygen enriched air and
CO2 is considered as one of the most favorable. In this case, it’s needed to enhance the boundary reaction
(C+CO2→2CO) using biomass chars produced under an N2/CO2/O2 atmosphere with CO2. If a CO-rich gas is yields
during the gasification process, so, the feed gas with desirable H2/CO ratio appropriate for the Fischer-Tropsch
synthesis reaction can be results without new H2 by promoting the water gas shift reaction (CO+H2O → CO2+H2) in
the downstream process at a lower temperature compared to that in the gasification step. In practice, CO2 gasification
of biomass char produced under an N2/CO2/O2 atmosphere has not been fully studied [92].

5.4. Hydrogen Gasification (insert the summary of last 3 researches related to this title done in 2016, 2017 and
2018)

Through this method of gasification, the biomass fuel is converted to gaseous fuels in the presence of hydrogen
under high pressures. This gasification process is critical that strict reaction conditions are maintained since the
majority of the products usually are in the gaseous phase. This process is undesirable, because of the strongly of control
which necessary as well as the fact that hydrogen must be readily obtainable. Weil et al. [93] used preheated hydrogen
mixed with peat at the entrance of fluidized bed gasifier. The reactor was operated as an entrained flow reactor in an

- 14 -
isothermal or a constant heat up mode. They reported that increase in temperature from 426C to 760C would increase
the carbon monoxide and hydrocarbon gases from 8% to 18% and 41% to 63%, respectively.

5.5. Carbon Dioxide Gasification (insert the summary of last three types of research related to this title done in
2016, 2017 and 2018)
Gasification is a thermo-chemical process, in which a solid or liquid fuel is converted at higher temperatures and in
the presence of gasification agent (air, H2O, O2 or CO2) to a mixture of combustible gases (H2, CO, CH4 and others
minor gases), accompanying gases (CO2, N2) and undesired components (tars, dust and others). Several attempts have
been made for the production of carbon monoxide from biomass with CO2 as the gasifying agent as follows.

Zhang, et al. [94] studied the effect of the pyrolysis mediums composition in the biomass fluidized bed gasification
process at 550C. their pyrolysis mediums were CO2, air, CO, H2, and CH4. The authors report that the liquid yield,
gas composition, and syngas heating value depend on the gasification agent composition. The gasification with CO2
at atmospheric pressure was producing less char than in the other agent. In case of CO2 as an agent, the CO2 in gas
yield decreased compared to the CO2 yield obtained in air. Furthermore, the CO2 led to the highest yield of acetic
acid compared to the other atmospheres. Zhang et al. illustrate these observations by two possible mechanisms; the
CO2 reacts with the active volatiles or with the char. newly, in their study on macro-algae pyrolysis/gasification, Kwon
et al. [95] observed that uses of CO2 in the pyrolysis process at 550C in a cylindrical gasifier minimized of the
generation of pyrolysis oil. They detected that, a decrease in the pyrolytic oil by the ratio of 24.3% and promoted
pyrolysis syngas production. This observation means that hydrocarbons yields from the pyrolysis process can be
decomposed in the presence of CO2. The same authors did another study about the pyrolysis of styrene butadiene
rubber [96] they reported that the CO2 increases C4 hydrocarbons cracking in addition to impeding the gas phase
addition reaction by which is formed benzene. Their pyrolysis experiments performed at 650 C in a free and CO2/ air
mixture. The authors showed that the number of condensable hydrocarbons decreased by the ratio of 30 to 50% when
adding CO2 to a modification of the syngas. The authors studied the effect of CO2 on a volatile matter during the
thermal dissolution of cellulose [96] at low and high heating rates of 500C/min. The authors reported that there is a
substantial increase in H2 and CO production in the presence of CO2. The authors performed experimental tests on
cellulose gasification in a cylindrical gasifier. The concentrations of H2, CO, and CH4 in the presence of CO2
increased by ratios of 4%, 10%, and 7%, respectively.

As the boundary reaction is only favorable at temperatures values higher than 700C, this increase of gaseous
compounds is related to the fact that CO2 likely accelerate the thermal cracking of volatile compounds release biomass
fuel. Approximately 67% decreased the condensable hydrocarbons collected during the experiments based on the mass
balance. The author evidenced that the primary influence of CO2 during biomass thermal degradation is perceived in
the gas phase was this latter found to participate in the hydrocarbons cracking leading to more permanent gases. The
same results were found using real biomass from corn stover. The authors did not observe noticeable differences in
the thermo-gravimetric analysis of the different biomasses degradation under CO2 and N2 respectively.

Kwon, et al. [96] performed macro-algae gasification tests with steam and variable concentrations of CO2 in a tubular
reactor with temperatures ranging from 600C to 1000C. The authors observed that the CO concentration increased by
a factor of 2 even at 600C and 700C with a CO2 level of 30% compared to a reference state of a CO2-free atmosphere.
The yield of C2 hydrocarbons also increased, and the amount of tar was reduced by 51.2% which can be directly
correlated with the gas yield augmentation. In their paper on CO2 as a carbon neutral fuel source via enhanced biomass
gasification, the authors studied the gasification of several kinds of wood, grasses, and agricultural residues with steam
and CO2 via thermo-gravimetric analysis and gas chromatography. The CO2 concentrations were varied between 0%
and 100% with steam as a co-reactant. The authors observed that, when only injecting 5% of CO2 with steam, CO
concentration increased by the ratio of 10% and H2 decreased by the rate of 3.3% at 900C.

Rising of CO2 from 5% to 50% resulted in continued CO increases and H2 decrease by the ratio of 3% at 900 C. A
high CO2 fraction in gasifying agent resulted in low H2 yield and high CO yield. The CO2 can potentially react in the
gas phase with hydrocarbons, such as methane, through dry reforming reaction. The CO2 can also react with hydrogen
atoms according to the reverse water gas shift reaction (RWGS). As well, in biomass gasifier; CO2 can react with the
carbon in the char form after the pyrolysis step through the heterogeneous boundary reaction. The primary results of
these reactions are the enhanced CO production. The authors’ conclusion, the introduction of CO2 decrease the energy
yield of the producer syngas, but on the other hand, reduce the preheating energy of gasifying agent resulting in a

- 15 -
higher thermal efficiency under testing conditions ranges. The most elevated thermal efficiency of the process without
O2 was 52% under N2 (40%), CO2 (60 %) at 850C. For the operation using O2, where the part of gasifying agent
preheating energy supplied by the partial biomass combustion the higher thermal efficiency was 60% under the CO2
(60%), O2 (8.3 %) and N2 (31.7%) atmosphere at 950C.

Beside excellent thermal efficiency, the authors discussed if the use of CO2 in biomass gasifies as an agent can provide
an N2-free syngas which is more suitable for the synthesis of liquid fuels and chemicals. M. Pohorel et al. [92] found
that the use of CO2 as an agent in the process of catalytic biomass gasification had a significant effect on the conversion
of biomass fuel into gaseous and on decreasing the tar yields. The authors performed their study on beech wood pellets
in a fluidized bed gasifier. The bed material was dolomitic limestone preheated to 500C with the gasification agent.
The gasification reactions were performed at 850C, respectively in H2O/O2, air/O2 and CO2/O2/air mixtures. The
higher cold gas efficiency was observed when gasifying biomass with a CO2 containing air.

6. Factors Affecting Gasification Performance (insert the summary of last 3 researches related to this title done
in 2016, 2017 and 2018)
Many variables appear effect on the gasification performance, syngas composition and tar yields, including bed
pressure, bed temperature, fluidization velocity, bed height, gasifying medium, ER, feed material moisture content,
biomass particle size, SFR and presence of catalysts. In the following, a brief discussion about more factors influencing
the gasification process.

6.1. Effect of Gasification Temperature


The gasifier bed temperature affects all of the chemical reactions occurs in the gasification process. In an experiment
conducted by Narvaez et al. [71], an increase in gasifier bed temperature from 700C to 850C, at a constant ER of 0.30,
the composition of the syngas produced was changed as; H2 concentration increased from 5% to 10%, CO increased
from 12% to 18%, CO2 decreased from 16% to 14%. Methane and C2H2 decreased minimally. The results from a
similar experiment conducted by Radmanesh et al. [98] validated the conclusions drawn from Narvaez et al. [71] in
that when the temperature of the gasifier bed is increased, and the ER is held constant, the concentrations of H2 and
CO in the syngas increased. From Petersen and Werther’s [99] study on a circulating fluidized bed gasifier using
sewage sludge as the fuel, a significant change in syngas composition was observed with an increase in bed
temperature from 530C to 730C at an ER of 0.30. The H2 to CO ratio doubled over this temperature range. In a sewage
sludge gasification experiment from de Andrés et al. [100], it was shown that higher bed temperatures favor hydrogen
production and that the concentration of CH4 increases slightly with increasing bed temperatures. Higher temperatures
produce more intense volatilizations and cracking reactions instead of creating more intense reforming reactions. Thus
at higher temperatures, there is an increase in H2 and light hydrocarbons in the syngas [100]. This pattern is shown
for temperature ranges of 750C-850C at constant ERs of 0.20, 0.30, and 0.40. Tar content also decreases as temperature
increases. These changes in composition affect the overall heating value and quality of the gas. An increase in bed
temperature, up to a point, increases the heating value and decreases the tar content which makes for a better quality
syngas. Note that increased bed temperature could have positive effects on the syngas quality (lower tar content) but
could also be detrimental to the amount of energy produced (lower heating value). Experimental literature has shown
that increases in temperature in sewage sludge gasifier-bed improve the heating value of the gas until the temperature
reaches 1470°F for a bubbling fluidized bed gasifier and 1340°F for a circulating fluidized bed gasifier [99, 94].

The FLETGAS system used in the laboratory experiment conducted by Gómez-Barea, et al. [101] consisted of a (1)
devolatilizer, a (2) reformer, and (3) a moving bed. The devolatilizer is where most of the volatile gases are released
from the biomass, and was operated at temperatures in the range of 700C-750C. These are high enough to release the
volatile gases from the sewage sludge but not high enough to cause any tar cracking, so a significant amount of tar
was released from the fuel in the devolatilizer. The tar and syngas were directed to another stage downstream of the
volatilization zone called the reformer. Oxygen-enriched air (40%O2) and high-temperature steam were blown into
the reformer to reduce the number of tars significantly by raising the temperature in the reformer to 1200C which
partially combusted the tars. The solids (char and ash) produced in the fluidized bed were then transferred to the third
stage, the moving bed. The reformer gases flowed into and transferred heat to the moving bed. The bed acted as a
catalytic filter in promoting tar decomposition reactions while steam introduced into the bed supported endothermic
char gasification reactions. Finally, the ashes exited the bottom of the moving bed containing very little carbon and
the syngas exited near the bottom of the moving bed containing negligible amounts of tar due to the two reduction

- 16 -
steps encountered in the process. The final stage of the system, the moving bed, cooled the ash and syngas streams
which increased the chemical energy of the syngas and thus the overall gasification efficiency.

6.3. Effect of Equivalence Ratio


The equivalence ratio (ER) is defined as; the ratio between the air flow rate introduced into the gasifier and
stoichiometric air flow rate that required for complete combustion/pyrolysis of the biomass fuel. It has also been stated
in the literature that it is one of the most significant operating conditions in biomass gasification process [71, 99].
O2 Supplied
ER = (2.15)
Theor. O2 Required
The O2 Supplied is the oxygen mass flow rate introduced into the gasifier (kg/hr.). Theoretical O2 required
is the notional amount of oxygen needed for complete combustion/pyrolysis of biomass (kg/hr.). ER dramatically
affects the syngas composition. As ER increases, the concentrations of combustible gases (H2, CO, CH4, and tars)
decrease while the levels of CO2 and H2O increase. Increases in ER provide a significant amount of Oxygen to the
gasifier which then goes to oxidize CO, H2, CH4, and tars. The methanation and oxidation reactions use O2 to oxidize
CH4 to CO and H2; hydrogen is oxidized to H2O. This trend can be seen in the literature over temperature ranges of
700-850C for both biomass and sewage sludge fuel [71, 99]. Recommendations in the literature for an optimal ER
vary based on feedstock and type of gasifier. In Petersen and Werther’s [99] experiment using sewage sludge as a fuel
and a circulating bed gasifier, the optimal ER was found to be 0.30.

Narvaez et al. [71] recommend values between 0.18 and 0.45 for the ER in their experiment using biomass as a fuel
and a bubbling fluidized bed gasifier. A lower ER is not practical because too much tar is produced and a higher ER
produces syngas with a little heating value. In Manyà, et al. [102] experiments with a dried sewage sludge fed fluidized
bed gasifier, an optimal ER was found at 0.35; the optimum qualification was determined by the highest concentration
of H2 in the syngas. Also, a high degree of combustion occurs at a higher value of ER; which supplies more amount
of air into the gasifier and promote char burning to produce more CO2 instead of combustible gases as H2, CO, and
CH4. Also, the ER increase results in a decrease in the low heating value (LHV) of the syngas because it hinders the
CH4 production other light hydrocarbons which have relatively large LHV. In additional to, at higher ER, extra
nitrogen provided by air dilutes the producer syngas which in turn results in its low energy content [102]. Studies have
shown that too small ER is also hostile to biomass gasification as it lowers the reaction temperature [71]. Therefore,
an optimum value for ER in biomass gasification exists in the range of 0.2-0.4 which differs according to various
operating parameters. Selection of the suitable ER is somehow depended on the producer gas subsequent application.
When the raw producer gas is going to be burnt in downstream furnaces, tar is not a severe issue. Also, the gas should
have a high heating value, and therefore the gasifier can be operated at the minimum ER of about 0.2. In the case of
temperatures lower than 850C, tar yield is high, and ER should be increased to about 0.3-0.4 to compensate such
adverse effects. Lv et al. [72] studied the effect of ER on gas yield and LHV. They varied the ER from 0.19 to 0.27
and realized that the variation of ER could be divided into two stages of 0.19-0.23 and 0.23-0.27. In the first stage, the
gas yield increased from 2.13 to 2.37 m3/kg, and the gas LHV rose from 8.82 to 8.84MJ/Nm3. It was observed that in
the second stage, the LHV and gas yield decreased due to the improvement of the oxidation reactions which also
reduced the concentration of CO, CH4, and CnHm and increased the CO2 level. So, the value of 0.23 was selected as
the optimum ER. In another set of experiments conducted by Narvaez et al. [71], ER was varied in the range of 0.25-
0.45 to find the optimum ER. It was observed that increasing the ER reduced the amount of H2, CO, CH4, and C2H2.
The maximum H2 concentration of 10% was obtained at ER of 0.26. They also realized that while the ER was
increased the tar content of the producer gas was gradually decreased and at ER of 0.45, minimum tar concentration
of 27g/m3 was achieved. They obtained LHV of 5.2-7MJ/m3 and 3.5-4.5MJ/m3 at ERs of 0.25 and 0.45, respectively.
It was also concluded that the gas yield was in a direct relationship with ER.

Similar trends were obtained by Li et al. [94] who investigated the co-gasification of biomass and coal while the ER
was in the range of 0.31-0.47. They also explained that as ER increased, more oxygen was introduced into the gasifier
which enhanced the combustion and increased the bed temperature from 948 to 1026C. Skoulou et al. [105] also
studied the effect of ER variation (0.2-0.4) as one of the most critical operation parameters on the quality of the
producer gas. They reported the favored concentration of CO at low ER of 0.2 and its hindered production at ER of
0.4 because of complete oxidation of carbon to CO2. Also, H2 production peaked at ER of 0.2. Lower heating value
of the producer gas was obtained at high ER which was due to the promotion of the oxidation reaction and dilution of
the producer gas with N2.

- 17 -
6.4. Effect of Steam to Fuel Ratio (insert the summary of last 3 researches related to this title done in 2016, 2017
and 2018)
Steam to fuel ratio (SFR) is defined as the flow rate of the steam fed into the gasifier divided by the biomass flow
rate is one of the critical process parameters involved in steam gasification [75]. An experimental study on biomass
air-steam gasification was conducted by Lv et al. [72]. They investigated the effect of SFR on the quality of the
producer gas in the range of 0-4.4. It was observed that the introduction of steam to the system improved the gas yield,
LHV and carbon conversion efficiency. They reported the SFR range of 1.35-4 as the optimum SFR in which the CO,
CH4 and C2H2 content of the producer syngas decreased, whereas the CO2 and H2 concentration gradually increased.
It was explained that in this SFR range, more steam reforming reactions of CO, CH4, and C2H2 occurred in the
presence of vapor which resulted in high concentrations of H2 and CO2. Over the optimum range, a decreasing trend
was observed in the syngas yield, LHV and carbon conversion efficiency due to the low reaction temperature affected
by low-temperature steam.

Qin et al. [104] examined the effect of SFR on tar formation and the resultant tar properties. In their experiments, SFR
was varied in the range of 0.49-2.66 at 900C. The results showed that, as the SFR was increased, the tar yield slowly
decreased from 3.87 to 1.71%. It was also concluded that high SFR values lower the aromaticity of the tar contents.
Another set of experiments was conducted by Gil et al. [106] who studied the effect of steam/oxygen gasification on
product concentrations. Their results showed steam to oxygen ratio, and steam/oxygen to biomass ratios was varied in
the range of 2-3 mol/mol and 0.6-1.6 kg/kg, respectively. The achieved results shown that the H2 content of the syngas
was in the field of 14-30% and decreased as the steam/oxygen to biomass ratio was increased, or the steam/oxygen
ratio was steadily reduced. As the O2 introduced into the system was improved, more H2 was combusted in the gasifier
and less was found in the discharge stream. A similar trend was observed for CO while changing the distinct ratios
and its concentration in the producer syngas was seen in the range of 30-50%. The tar content of the syngas naturally
reduced to less than 10 gram /m3 as the steam/oxygen to biomass ratio was increased to 1.0-1.1kg/kg. The char produce
also decreased to 10% while the gasifying agent to biomass ratio was expanded to the values higher than 1.0.

As mentioned previous, steam gasification can provide a gas stream with a high content of H2, but the concentration
of the unattractive products, for example, CO2 is also increased. Therefore, to improve the efficiency of the steam
gasification process, considerable efforts have been dedicated to the production of producer gas with a high yield of
H2 with the instantaneous capture of CO2.

6.5. Effect of Biomass Size (insert the summary of last 3 researches related to this title done in 2016, 2017 and
2018)
It has been known that small particle size biomass significantly increases the overall energy efficiency of the
gasification process, but it also increases the gasification plant cost. It has been estimated that for a 5-10MW
gasification plant, about 10% of the output energy is required for the biomass particle size reduction [107]. On the
other side, an increase in biomass particle size reduces the pretreatment costs, but the devolatilization time increases,
and thus for a defined throughput the gasifier size increases [107]. Hence, a balance should be considered while
investigating the effect of biomass particle size on the gasification efficiency. Lv et al. [39] studied the impact of
biomass size on the quality of the producer syngas in four ranges as of 0.6-0.9, 0.45-0.6, 0.3-0.45 and 0.2-0.3mm.
They detected that small particle size biomass created more amounts of CH4, CO and C2H4 and less CO2 in
comparison to large particles. Consequently, the producer syngas yield, LHV, and carbon conversion were enhanced
as the biomass particle size decreased. It was clarified that small biomass particles contribute to the large surface area
and high heating rate which in turn produce more light gases and less char and condensate. Thus, the flow rate and
composition of the producer syngas improved while using the small particle biomass.

6.6. Effect of Bed Materials (Catalyst) (insert the summary of last 3 researches related to this title done in 2016,
2017 and 2018)
Bed materials are of great significance in fluidized bed gasifiers. They turn as heat transfer medium, but their crucial
role involves in tar cracking which avoids complex downstream tar removal process [108]. The presence of a catalyst
in the bed material during biomass gasification promotes several chemical reactions which influence the composition
and heating value of the producer gas. It also reduces the tar yield and prevents stable agglomeration tendency of the
bed [89]. The catalytic reforming reactions through which tar is converted into valuable gaseous compounds are
summarized as follows [89]:

- 18 -
CnHm+ nH2O ↔ (n+0.5 m) H2+ n CO
CnHm+ nCO2 ↔ 0.5m H2+2n CO
CnHm+ (0.5n+0.25m) O2 ↔ 0.5m H2O+ n CO
CnHm↔ 0.5m H2+ n C
The three main groups of catalysts are implemented to eliminate tar from the producer syngas [108]: (1)
natural catalysts such as dolomite and olivine; (2) alkali-based catalysts such as (Li, Na, K, Rb, Cs and Fr) and (3)
metal-based catalyst such as nickel catalysts. Dolomite is the most usually used catalyst which effectively removes
heavy hydrocarbons from the syngas stream [89,108,109]. It also decreases accumulation in fluidized bed while using
biomass with high alkali content. But, the undesired property of dolomite is its rapid calcination in the gasifier which
subsequently results in a syngas with high particulate. Alkali-based catalysts (Li, Na, K, Rb, Cs, and Fr) are capable
of increasing the gasification rate and decreasing the tar content of the producer syngas. However, effort in recovery,
high cost, and accumulation at high temperatures are some of the difficulties of the alkali-based catalysts [109]. Metal-
based catalysts are also significantly effective in eliminating tar and increase the superiority of the producer syngas.
The main challenges associated with this type of catalysts are carbon deposition and nickel particle growth, which
cause catalyst deactivation [108].

Asadullah et al. [108] studied the performance of the heterogeneous catalyst of Rh/CeO2/SiO2 in fluidized bed
gasification to that of dolomite, steam reforming catalyst and inert bed materials whereas the ER was set at 0.31 and
the bed temperature within the range of 823K-973K. It was detected that the tar content of the producer was utterly
negligible while using Rh/CeO2/SiO2 as the bed material. However, the tar concentration at about of 113 and
139gram/m3 was achieved with dolomite and inert bed materials, respectively. It was also concluded that in the case
of Rh/CeO2/SiO2 catalysts the efficiency of cold gas was about 71% more than others cases. Also, little char and coke
were observed in the experiments with the Rh/CeO2/SiO2 as catalysts. Skoulou et al. [105] used quartz sand and
olivine as bed materials in a bubbling fluidized bed gasifier at ER of 0.2-0.4 and bed temperature as 750C-850C. They
reported that, while quartz sand is an economy and abundant element, it produced severe de-fluidization due to its
tendency to tar formation at temperatures below 800C. They replaced quartz sand with olivine and observed that, at
low gasification temperature of 750C and ER of 0.2, components of tar were pyrolysis and released H2 and CO, under
the catalytic effect of iron-based olivine.

Li et al. [94] examined the influence of bed material on tar elimination efficiency in a circulating fluidized bed. They
used silica sand and a commercial Ni-alumina catalyst as bed material. At the bed temperature of 800C, the quantity
of tar reduced from 0.4 gram/m3 to 0.15 mg/m3, as silica sand was substituted with Ni-alumina catalyst.

7. Modeling of Biomass Gasification Process (insert the summary of last 3 researches related to this title done
in 2016, 2017 and 2018)
The primary goals of models are to study the thermo-chemical processes during the gasification of the biomass and
to evaluate the influence of the primary input variables, such as steam to fuel ratio, equivalence ratio and gasification
temperature on syngas composition and its heating value. In the present section, the literature survey has been carried
out for mathematical models, ASPEN PLUS models and neural network models.

7.1. Simulation and Mathematical Modeling Studies


Mathematical simulation is one of the most critical aspects of research and development work such as the
development of gasification technology. Though it may not provide a very accurate prediction of systems performance,
it may provide qualitative guidance on the effect of design, input variables and operating conditions. Moreover,
modeling may provide a less expensive means of evaluating the benefits and the associated risk in the real-time
scenario [110]. Gasifier simulation models may be classified into thermodynamic equilibrium model, kinetic model,
computational fluid dynamics (CFD) model and artificial neural network. All these models approach different methods
to assess the prediction of the one’s model and have separate utility and limitations. However, modeling of gasification
using different approach may consider the Boudouard reaction following reaction as necessary gasification reaction
[112, 113].

Boudouard reaction (R1) C+CO2  2CO (2.1)


Water-gas reaction (R2) C+H2O H2+CO (2.2)
Methane formation (R3) C+2H2  CH4 (2.3)

- 19 -
Steam reforming reaction (R4): CH4 +2H2O  4H2 + CO2 (2.4)
7.1.1. Thermodynamic Equilibrium Model [114]
Thermodynamic equilibrium models are based on the chemical and thermodynamic equilibrium, which is
determined by implication of equilibrium constants and minimization of Gibbs free energy. At chemical equilibrium,
the system is considered to be at its most stable composition, which means the entropy of a system is maximized,
while its Gibbs free energy is minimized. Though chemical or thermodynamic equilibrium may not be reached within
the gasifier, equilibrium models provide a designer with a reasonable prediction for the final composition and monitor
the process parameter like temperature [112]. Some significant assumptions of thermodynamic equilibrium are
presented below:

1. The reactor is considered as zero-dimensional [115]


2. There is perfect mixing of materials and uniform temperature in the gasifier although different hydrodynamics are
observed in practice [112]
3. The reaction rates are fast enough and residence time is long enough to reach the equilibrium state [112].

Equilibrium models are independent of gasifier design and cannot predict the influence of hydrodynamics or
geometric parameters like fluidizing velocity, design variables (gasifier height). However, these models are quite
convenient to study the impact of fuel and the process parameters and can predict the temperature of the system [111].
Thermodynamic equilibrium models can be approached by either stoichiometric or non-stoichiometric methods as
follows.

7.1.1.1. Stoichiometric Equilibrium Models [117]


Stoichiometric equilibrium models incorporate the thermodynamic and chemical equilibrium of chemical reactions
and the species involved. The model can be designed either for a global gasification reaction or can be divided into
sub-model for drying, pyrolysis, oxidation, and reduction.

7.1.1.1.1. Single step stoichiometric equilibrium model [116]


This model embodies the several complex reactions of gasification into one general reaction as mentioned in Eq.
(2.5). It assumes that one mole of biomass CHxOy, based on a single atom of carbon that is being gasified with w
mole of water/steam in the presence of a mole of air as the following general equation.

CHx Oy +wH2O+a  O2 +3.76N2  =n cC+n H2H2 +n coCO+n Co2 CO2 +n CH4CH4 +n H2OH2O+3.76aN2
(2.5)

In the above equation, w and a, are the variables and changed in order to get desired amount of product. There are
n ,n n n n H2O
six unknowns are nc, H2 co , CO2 , CH4 and . Based on stoichiometric balance of carbon, hydrogen and
oxygen, following equations are obtained:
n c +n co +n Co2 +n CH4 =1
Carbon balance: (2.6)
2n H2 +4n CH4 +2n H2O =2w+x
Hydrogen balance: (2.7)
n +2n
CO +n
CO2 H2O =w+2a+y
Oxygen balance: (2.8)
As Boudouard reaction, water-gas reaction, methane formation and steam reforming reaction are considered
as the major reaction of gasification, the equilibrium constants (keq) for reactions R1, R2 R3 and R4 are given as
Perry et al., [142].
n 2 CO
k eq1 =
n CO2 (2.9)
n H2 .n CO
k eq2 =
n H2 O
(2.10)

- 20 -
n CH4
k eq3 =
n 2 H2
(2.11)
n H2 .n CO2
K eq4 =
n CO .n H2 O
(2.12)
If the gasification process is assumed to be adiabatic, then the energy balance of the gasification reaction results to a
new set of equation, which can determine the final temperature of the system [116].

 n h
i
i
0
f,i +ΔH T298 
i, Reactant
=  n i  h 0 f,i +ΔH T298 
i
i, Product loss
(2.13)
Modifying Eq. (2.13) on the basis of Eq. (2.5), we get:
 0

h 0f, wood +w h 0f, H2O1 +h vap +ah f,O 2
+3.76ah f,0 N2 =n c .h f,c
0 0
+n H2 h f,H 2
0
+n CO h f,CO +n CO2 +n CH4 h f,0 CH 4 +n H2O h 0f,H2O
+3.76ah 0f, N2 +ΔT(n C Cp,C +n H2 Cp,H2 +n CO Cp,CO +n CO2 Cp,CO2 +n CH 4 Cp, CH 4 +n H 2O Cp, H 2O +3.76aCp, N2 )
(2.14)
In this equation,h0 f , Cp,C , hvap represents heat of formation of corresponding chemical species, specific heat capacity
and enthalpy of vaporization of water respectively and ΔT=Tgasification-Tambient refers to temperature difference between
the gasification temperature and the ambient or the initial temperature of biomass feedstock [110,114]. Single step
stoichiometric equilibrium model may be formulated by the application of the chemical equilibrium state and the
reaction stoichiometric condition.

7.1.1.1.2. Sub-models for stoichiometric equilibrium model [117]


This model incorporates modeling of separate sub-model for drying, pyrolysis, oxidation and reduction, the
output from one sub-model becomes input for the successive sub-model. This model has more utility than the single
step stoichiometric equilibrium model as the composition and temperature at different zone can be assessed with the
aid of sub-model. Several combinations of sub-models can be achieved and can be selected as per the requirement of
the model and its feasibility. For the sake of convenience and clarity of sub- model, sub-models for drying and
pyrolysis, oxidation and reduction zone have been proposed for the current paper. However, the modeling approach
follows similar principle as that of single step stoichiometric equilibrium model regarding the mathematical
formulation. One of the uncertainties of such sub-model lies in their assumption for final product. For example, the
assumptions implied in pyrolysis sub-model indicate that that the product composition mainly includes CO, CO 2 H2
,H2O, CH4 and tar with higher concentration of lighter component as in Eq. (2.14) [118]. The pyrolysis products
undergo partial oxidation in presence of non-stoichiometric oxygen supply, and the reaction in oxidation sub-model
may be proposed as in Eq. (2.15) [117, 119].

7.1.1.2. Non-stoichiometric Equilibrium Model [120]


The non-stoichiometric equilibrium model is solely based on minimizing Gibbs free energy of the system
and there is not any specification for particular reaction mechanisms. However, moisture content and elemental
composition of the feed is needed which can be obtained from the ultimate analysis data of feed. Therefore, this method
is particularly suitable for fuels like biomass whose exact chemical formula is not distinctly known [110]. The Gibbs
free energy, Gtotal for the gasification product which consists of N species (i=5…N) is represented as in Eq. (2.19)
[110].
N N  n 
G total =  n i ΔG 0f,i + n i RTln  i 
 n 
i=1 i=1  i 
(2.19)
0
Where, ni ∆Gf,i is the standard Gibbs energy of i species, R is gas constant. The solution of Eq. (2.19) for
unknown values of ni is approached to minimize Gtotal of the overall reaction considering the overall mass balance,
Though, non-stoichiometric equilibrium model does not specify the reaction path, type or chemical formula of the
fuel, the amount of total carbon obtained from the ultimate analysis must be equal to sum of total of all carbon
distributed among the gas mixtures (Eq. (2.20)) [120].

- 21 -
N

 =a
i=1
n
i,j i
(2.20)
Where a i,j is the number of atoms of the j element and Aj is the total number of atoms of jth element in
reaction mixture. The objective of this approach is to find the values of ni such that the Gtotal will be minimum.
Lagrange multiplier [120]. Thus, the Lagrange function (L) can be defined as:
K
 N 
L=G total - λj   a i,jn i -Ai 
i1  i=1  (2.21)
Where λ is Lagrangian multipliers. The equilibrium is achieved when the partial derivatives of Lagrange function
are zero. i.e.,
 L 
  =0
 n i  (2.22)
Dividing Eq. (2.21) by RT and substituting the value of G total from Eq. (2.19,) then taking its partial derivate results
to Eq.

 L  ΔG 0f,i N  ni  1 K  N 
+  ln
 =
 n i  RT i=1
 + RT  λj  a i,jn i =0
 i1 
 n total  j=1

(2.23)
The standard Gibbs free energy of each chemical species can be obtained by subtracting the standard enthalpy form
the standard entropy multiplied by a specific temperature of the system as in Eq. (2.24) [110].
ΔG0j,i =ΔHf,i0 +TΔSf,i0
(2.24)
Where ∆S̅f,i
0
is the standard entropy of i species According to first law of thermodynamics, the energy balance of the
non- stoichiometric equilibrium model can be achieved by Eq. (2.25) [110, 120].
 n r H r0 T r   Q loss   n p H pt0 T pt   H
r  reactant pt  product
(2.25)
Thus, the final compositions of the product gas can be determined via non-stoichiometric equilibrium
approach. Moreover, this model gives the utility to examine the effect on product gas composition and temperature
by changing the moisture content and biomass feed. However, such models have plenty of limitations. Srinivas et al.
[122] did thermodynamic modeling studies to predict the performance characteristics of a rice husk based integrated
gasification combined cycle plant at the variable operating conditions of gasifier. They applied for wet fuel (fuel
with moisture) for predicting the gas composition, gas generation per kg of fuel, plant efficiency and power
generation capacity, and NOx and CO2 emissions.

7.2. ASPEN PLUS Gasification Models

ASPEN Plus is a problem-oriented input program that is used to facilitate the calculation of physical,
chemical and biological processes. It can be used to describe processes involving solids in addition to vapor and liquid
streams. ASPEN Plus makes the model creation and updating easier since small sections of complex and integrated
the systems can be created and tested as separate modules before they are combined. This process simulator is equipped
with a large property data bank containing the various stream properties required to model the material streams in a
gasification plant, with an allowance for the addition of in-house property data.

Mathieu and Dubuisson [123] modeled wood gasification in a fluidized bed using ASPEN Plus. The model was based
on the minimization of the Gibbs free energy, and the process was uncoupled in pyrolysis, combustion, Boudouard
reaction and gasification. They performed a sensitivity analysis and concluded that there is a critical air temperature
above which preheating is no longer efficient, that there is an optimum oxygen factor, that the oxygen enrichment of
air plays an active role under a particular value, and that the operating pressure has only a slightly positive effect on
process efficiency. Mitta et al. [124] modeled a fluidized-bed type gasification plant with air and steam using ASPEN
Plus and validated their results with the gasification pilot plant located at the Chemical Engineering Department of the

- 22 -
Technical University of Catalonia. Their gasification model was divided into three different stages: drying,
devolatilization pyrolysis, and gasification-combustion. Nikoo and Mahinpey [125] developed a model capable of
predicting the steady-state performance of an atmospheric fluidized-ban ed gasifier by considering the hydrodynamic
and reaction kinetics simultaneously. They used four ASPEN Plus reactor models and external FORTRAN subroutines
for hydrodynamics and kinetics nested to simulate the gasification process.

The ASPEN Plus yield reactor, “RYIELD,” was used to affect the decomposition of the feed. They validated their
model using different sets of operating conditions for a lab-scale pine gasifier with air and steam. Paviet et al. [128]
proposed, a thermochemical equilibrium model using ASPEN Plus for downdraft and staged gasifiers. They claimed
that proposed model is easy to build and predicted with accuracy the composition of the flaming pyrolysis gas and the
producer gas. The reaction temperature is the parameter that controls the whole gasification process. It influences the
final producer gas directly. The models developed in this work can be used as input parameters in the design of a
gasifier. Concentrations given by the flaming pyrolysis gas model can be used as input parameters to the char
gasification model. Levels provided by the producer gas model used as input data to the combustion model of an SI
engine, to design or to predict the overall performance, of a gasification unit. Naveed et al. [131] developed steady-
state simulation model for gasification using ASPEN Plus. The model can be used as a predictive tool for optimization
of the gasifier performance. The authors of the study modeled the gasifier in three stages. In first stage moisture content
of biomass, the feed is reduced. In second stage biomass is decomposed into its elements by specifying yield
distribution. In third stage gasification reactions have been modeled using Gibbs free energy minimization approach.
The simulation results were compared with the experimental results obtained through hybrid biomass gasifier. They
concluded that higher temperature improves gasifier performance. It increases production of carbon monoxide and
hydrogen in syngas which ultimately results in higher heating value and gas conversion efficiency. Equivalence ratio
controls the production of syngas by controlling carbon conversion of fuel and extent of gasification reactions.

Damaris et al. [132] carried out an assessment of combined heat and power (CHP) biomass bubbling fluidized bed
gasification unit coupled with an internal combustion engine (ICE) by using a comprehensive mathematical model
based on the ASPEN PLUS process simulator. The proposed model is based on a combination of modules that ASPEN
PLUS simulator provides representing the three steps of gasification process (drying, pyrolysis, and oxidation), gas
cleaning and the ICE. The model is also based on mass and energy balances and reaction kinetics.

7.3. Neural Network Models


Since the invention of the digital computer, researchers attempted to create machines which directly interact with
the real world without their intervention. In this sense, the Artificial Intelligence, in general, and particularly the
Artificial Neural Networks (ANN’s) represent an alternative for endowing to the computers one of the characteristics
that make the difference between humans and live beings, the intelligence. An artificial neural network is an abstract
of a real nervous system, and its study corresponds to growing interdisciplinary fields which consider the systems as
adaptive, distributed and mostly nonlinear, three of elements found in the real applications. Artificial neural networks
(ANN) have been extensively used in the fields of pattern recognition, signal processing, biomedical instrumentation,
function approximation and process simulation. The Neural Networks package supports different types of training or
learning algorithms. Pavlas et al. [134] have proposed a biomass gasification system for complex design interactions
as many streams requiring heating and cooling in the energy recovery. The conceptual understanding gained from the
case study provides systematic design guidelines for further process development and industrial implementation in
practice.

Sipöcz et al. [135] present the development and validation of an ANN model of a CO2 capture plant. An evaluation
of the concept is made of the usefulness of the ANN model as well as a discussion of its feasibility for further
integration into a conventional heat and mass balance programme. It is shown that the trained ANN model can
reproduce the results of a rigorous process simulator in a fraction of the simulation time. A multilayer feed-forward
form of Artificial Neural Network was used to capture and model the non-linear relationship between inputs and
outputs of the CO2 capture process. They concluded that Artificial Neural Networks are found to be useful tools for
predicting complex processes such as CO2 capture processes, which, when simulating the closed process network,
yields a challenging solution pathway which is computationally demanding and challenging to reproduce using
traditional methods. The average value of the errors for the prediction of specific re-boiler duty is well below 0.2%,
and the maximum error does not exceed 3.1%. The prediction of solvent rich load and amount CO2 captured are even
better, with the maximum error below 2.8% and 0.17% respectively. Puig et al. [136] applied Artificial neural networks

- 23 -
(ANNs) used for modeling biomass gasification process in fluidized bed reactors. Two architectures of ANNs models
are presented; one for circulating fluidized bed gasifiers (CFB) and the other for bubbling fluidized bed gasifiers
(BFB). Both models determine the producer gas composition (CO, CO2, H2, and CH4) and gas yield. Published
experimental data from other authors has been used to train the ANNs. The obtained results show that the percentage
composition of the primary four gas species in producer gas (CO, CO2, H2, and CH4) and producer gas yield for a
biomass fluidized bed gasifier can be successfully predicted by applying neural networks. ANNs models use in the
input layer the biomass composition and few operating parameters, two neurons in the hidden layer and the
backpropagation algorithm. The results obtained by these ANNs show high agreement with published experimental
data used R2 > 0.98.

8. Summary of Literature Review


In this chapter, an overview of biomass gasification and research activities around the world, as well as literature
study regarding the central research questions related to biomass gasification has been performed. The literature review
includes fluidized bed and their type, an overview of biomass used in gasification, tar formation and tar reduction and
tar measuring techniques, the effect of the agent on gasification and factors affect of gasifiers performance and
different gasification modeling methods. Significant topics of biomass gasification are outlined to identify
opportunities and potential challenges facing biomass gasification. Particular attention is paid to review the different
approaches to model biomass gasification process, i.e., thermodynamic equilibrium models and other models.
Observations and gaps have been listed at the end along with the objectives for the present work. Several important
points can be concluded from this literature study:

1. The biomass was observed to be sustainable energy resource for agriculture dependent economic countries.
2. Thermo-chemical conversion or gasification is a perspective and viable route to make use of biomass in the energy
generation in a decentralized manner.
3. Performance of gasification is dependent on reaction zone temperature, type of fuel and fuel prosperities,
equivalence ratio, steam to fuel ratio and agent type.
4. The gasification agent, strongly affecting on tar yield and syngas quality.
5. A few studies have been performed on a carbon dioxide and pure oxygen as gasification agent, also limited work is
reported on sawdust as biomass fuel.
6. More studies were conducted on tar minimization by different researchers. However, the formation and distribution
behavior of tar produced from the gasification of various biomass fuels still lacks in the literature. Despite the fact that
an experimental study of tar yield during biomass gasification is critical, the process could be time-consuming as well
as a challenge due to limitations and availabilities of tar measuring techniques. Thus, the thermodynamic equilibrium
simulations of the distribution performance of the variation of tar yields during biomass gasification will be helpful to
optimize gasifier operating conditions gas cleaning systems.
7. Using primary tar reduction methods is very attractive since working conditions have a significant influence tar
yield and formation. Therefore, to investigate the fate of tar during biomass gasification, tar measurement and analysis
is primarily required.
8. There is a significant variation among kinetic parameters derived from combustion and gasification of different
biomass chars reported by various researchers depending on biomass fuel types and fuel properties.
9. A few researchers have developed thermodynamic based equilibrium and ANN-based models to predict the gas
composition and gasifier performance.
10. None of the researchers reported the development of GA found ANN models to predict producer gas composition
from downdraft gasifier.

Conclusion
Research and development on new process of gasification is directed towards production of hydrogen-rich gas from
biomass the parameters has affected on the gasification of biomass such as (temperature, gasifying agent/biomass
ratio, pressure) and of the materials, type of biomass, type of gasifier) on the performance of the gasification system
was clarified which allow to increase yield of hydrogen gas. Critical parameters for success are the feedstock properties
and the feedstock pretreatment. For the third world, the use of a robust and straightforward technology represented by
gasification can assist the development of the rural economies by providing the electricity produced from local sources
of biomass.

- 24 -
References

[32] Elif Kirtay “Recencnt advance in production of hydrogen from biomass” Energy conversion and management, vol.52,
2011, pp.1778-1789.
[33] Maria Aznar P., Jose Corella, Jesus Delgado and Joaquin Lahoz, “Improved steam gasification of lingo-cellulosic
residues in a fluidized bed with commercial steam reforming catalysts” Ind. Eng. Chem. Res. vol. 32,1993.
[34] Lian, C.K. and M.E. Findley. “Air blown wood gasification in large fluidized bed reactor” Ind. Eng. Chem. Process
Dec, vol.21, 1982, pp.699-701.
[35] FAO. Food and Agriculture Organization of the United Nation, 1986.
[36] Raman, K.P., W.P. Walawender, L.T. Fan and C.C. Chang. “Mathematical model for fluid bed gasification of
biomass material” Application to feedlot manure. Ind. Eng. Chem. Process Des, vol.20, 1981, pp.686-692.
[37] Keiichi T., M. Asadullah, T. K. Kunmori, “Syngas production by biomass gasification using Rh/CeO2/SiO2 catalysts
and fluidized bed reactor” Catalysis Today. vol.89, 2004. pp.389-403.
[38] Khater E. M. K., N. N. El-Ibiary, I. A. Khatab and M. A. Hamad “Gasification of rice hulls” Biomass and Bioenergy,
vol.3,1992, pp.329-333.
[39] Midilli A., M. Dogru, Gr. Howarth, T., Ayhan, “Hydrogen production from hazelnut shell by applying air blown
downdraft gasification technique” International Journal of hydrogen energy, vol.26, 2001, pp.29-37.
[40] Abdul Salam P., S.C.Bhattacharya “A comparative study of char coal gasification in two types of spouted bed
reactors” Energy, vol.31, 2006, pp.228-243.
[41] Chen, C.C. and M.H. Rei. “Gasification of rice husk” Presented at Bio-Energy, World Congress and Exposition,
Atlanta, GA, vol.80, April 21, 1980.
[42] Sadaka, S. S., A. E. Ghaly and M. A. Sabbah. “Two phase biomass air-steam gasification model for fluidized bed
reactor: Part I, II, III” Biomass and Bioenergy, vol.22, 2002, pp. 439-487.
[43] Risnes, H., J. Fjellerup, U. Henriksen, A. Moilanen, P. Norby, K. Papadakis, D. Posselt, L. Sorensen. “Calcium
addition in straw gasification” Fuel, vol.82, 2003, pp. 641-651.
[44] Walawender, W.P., S. Ganesan and L.T. Fan. “Steam gasification of manure in a fluid bed: Influence of limestone as
a bed additive” IGT Symposium on Energy from Biomass and Wastes V, Lake Buena Vista, FA, Jan. 26-30, 1981.
[45] Yin, X., C. Zhi, S. Zheng and Y. Chen. “Design and operation of a CFB gasification and power generation system
for rice husk Biomass” Bioenergy, vol.23, 2002, pp.181-187.
[46] Singh, S.K., W.P. Walawender, L.T. Fan and W.A. Geyer. “Steam gasification of cotton wood (branches) in a
fluidized bed” Wood Fiber Sci., vol.18, 1986, pp.327-344.
[47] Groves, J.D., J.D. Craig, W.A. Le Pori and R. G. Anthony. “Fluidized bed gasification of cotton gin waste” ASAE,
1979, pp.79-4547.
[48] Epstein, E., H. Kosstrin and J. Alpert. “Potential energy production in rural communities from biomass and wastes
using a fluidized-bed pyrolysis system” IGT Symposium on Energy from Biomass and Wastes, Washington, D. C.,
Aug. 14-18, 1978.
[49] Wander, P., C. Altafini and R. Barreto. “Assessment of a small sawdust gasification unit” Biomass and Bioenergy,
vol.27, 2004, pp.467-476.
[50] Hoveland, A.D., W.P. Walawender, L.T. Fan and F.S. Lai. “Steam gasification of grain dust in a fluidized bed reactor”
Transactions of ASAE, vol.25, 1982, pp.1074-1080.
[51] Sweeten, J., K. Annamalai, B. Thien, L. McDonald. “Co-firing of coal and cattle feedlot biomass (FB) fuels. Part I.
Feedlot biomass (cattle manure) fuel quality and characteristics” Fuel, vol.82, 2003, pp.1167-1182.
[52] Walawender, W.P., D.A. Hoveland and L.T. Fan. “Steam gasification of Alpha- cellulose in a fluid bed reactor”
presented at Fundamentals of Thermochemical Biomass Conversion Conference, Estes Park, CO, 1982.
[53] Halligan, J.E., K.L. Herzog and H. W. Parker. “Synthesis gas from bovine wastes” Ind. Eng. Chem. Process Des,
vol.14, 1971, pp.64-69.
[54] Evans, R., Milne, T. “Chemistry of tar formation and maturation in the thermochemical conversion of biomass” Fuel
and Energy, vol.39, 1998, pp.197-198.
[55] Li, C., Suzuki, K. “Tar property, analysis, reforming mechanism and model for biomass gasification-An overview”
Renewable and Sustainable Energy Reviews, vol.13, 2009, pp.594-604.
[56] Li, J., Wang, Z., Yang, X., Hu, L., Liu, Y., Wang, C. “Evaluate the pyrolysis pathway of glycine and glycylglycine
by TG-FTIR” Analytical and Applied Pyrolysis, vol.80, 2007, pp. 247-253.
[57] Milne, T., Abatzoglou, N., Evans, R. “Biomass gasifier “tars”: their nature, formation, and conversion” National
Renewable Energy Laboratory, NREL/TP-570-25357, 1998.

- 25 -
[58] Boerrigter, H., Calis, H.P., Slort, D.J., Bodenstaff, H., Kaandorp, A.J., den Uil, H., Rabou, L.P.L.M. “Gas cleaning
for integrated biomass gasification (BG) and Fischer-Tropsch (FT) systems; experimental demonstration of two BG-
FT systems” ECN, 2004.
[59] Bui, T., Loof, R., Bhattacharya, S. “Multi-stage reactor for thermal gasification of wood” Energy, vol.19, 1994,
pp.397-404.
[60] Devi, L., Ptasinski, K.J., Janssen, F.J.J.G. “A review of the primary measures for tar elimination in biomass
gasification processes” Biomass and Bioenergy, vol.24, 2003, pp.125-140.
[61] Campoy, M., Go mez-Barea, A., Fuentes-Cano, D., Ollero, P. “Tar Reduction by Primary Measures in an Autothermal
Air-Blown Fluidized Bed Biomass Gasifier” Industrial & engineering chemistry research, vol.49, 2010, pp.11294-
11301.
[62] Corella, J., Aznar, M.P., Gil, J., Caballero, M.A. “Biomass gasification in fluidized bed: where to locate the dolomite
to improve gasification” Energy & Fuels, vol.13, 1999, pp.1122-1127.
[63] Sutton, D., Kelleher, B., Ross, J.R.H. “Review of literature on catalysts for biomass gasification” Fuel Processing
Technology, vol.73, 2001, pp. 155-173.
[64] Neeft, J., Knoef, H., Nederland, S.E.C., Group, B.B.T. “Behavior of Tar in Biomass Gasification Systems: Tar
Related Problems and Their Solutions” Novem, 1999.
[65] Kinoshita, C., Wang, Y., Zhou, J. “Tar formation under different biomass gasification conditions” Analytical and
Applied Pyrolysis, vol.29, 1994, pp.169-181.
[66] Li, X., Grace, J., Lim, C., Watkinson, A., Chen, H., Kim, J. “Biomass gasification in a circulating fluidized bed”
Biomass and Bioenergy, vol.26, 2004, pp.171-193.
[67] Van Paasen, S.V.B., Kiel, J.H.A. “Tar formation in fluidized-bed gasification-impact of gasifier operating conditions”
The 2nd world conference and technology exhibition on biomass for energy, industry and climate protection, ECN-
RX-04-037, 2004.
[68] Kurkela, E., Ståhlberg, P., Laatikainen, J., Simell, P. “Development of simplified IGCC-processes for biofuels:
supporting gasification research at VTT” Bioresource technology, vol.46, 1993, pp.37-47.
[69] Simell, P.A., Leppälahti, J.K. “Catalytic purification of tarry fuel gas with carbonate rocks and ferrous materials”
Fuel, vol.71, 1992, pp.211-218.
[70] D.L Brink, Goldstein, I.S., “Organic Chemicals from Biomass” CRC press, Florida, Chapter 4, 1981, pp.45.
[71] Narvaez, I., Orio, A., Aznar, M.P., 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” Industrial & engineering chemistry
research, vol.35, 1996, pp.2110-2120.
[72] Lv P., Z. Yuan, C. Wu, L. Ma, Y. Chen and N. Tsubaki “Bio-syngas production from biomass catalytic gasification”
Energy Conversion and Management, 2007, pp.1132-1139.
[73] Herguido, J., Corella, J., Gonzalez-Saiz, J. “Steam gasification of lignocellulosic residues in a fluidized bed at a small
pilot scale. Effect of the type of feedstock” Industrial & engineering chemistry research, vol.31, 1992, pp.1274-1282.
[74] Aznar, M.P., Caballero, M.A., Gil, J., Martin, J.A., Corella, J. “Commercial steam reforming catalysts to improve
biomass gasification with steam-oxygen mixtures. Part 2. Catalytic tar removal” Industrial & engineering chemistry
research, vol.37, 1998, pp. 2668-2680.
[75] Ergudenler, A. “Gasification of Wheat Straw in a Dual-Distributor Type Fluidized Bed Reactor” Unpublished Ph.D.
Thesis. Technical University of Nova Scotia. Nova Scotia. Canada, 1993.
[76] Cao, Y., Y.Wang, J. Rieley and W. “A novel biomass air gasification process for producing tar-free higher heating
value fuel gas” Fuel Processing Technology, vol.87, 2006, pp.343-353.
[77] Boateng, A.A., W.P. Walawender, L.T. Fan and C.S. Chee. “Fluidized bed gasification of rice hull” Bioresource
Technology, vol.40, 1992, pp.235-239.
[78] Slapak, M. J., J. M. van Kasteren, A.A.Drinkenburg. “Design of a process for steam gasification of PVC waste”
Resources, Conservation and Recycling, vol.30, 2000, pp.81-93.
[79] Mermoud, F., F. Golfier, S. Salvador,Van de Steene and J Dirion. “Experimental and numerical study of steam
gasification of a single charcoal particle” Combustion and Flame, vol.145, 2006, pp.59-79.
[80] Wei L, Xu S, Zhang L, Liu C, Zhu H, Liu S. “Steam gasification of biomass for hydrogen-rich gas in a free-fall
reactor” Int. J Hydrogen Energy, vol. 32, 2007, pp.24-31.
[81] Bailie, R.C. “Hessleman Gas Generator Testing” Solar Energy Research Institute. Contract No. AH-8-1077-1, 1979.
[82] Tillman, D.A. “Biomass Combustion. Biomass: Regenerable Energy” Hall, D.O. and Overened, R. P. (eds). John
Wiley and Sons, 1987, pp.203-219.
[83] Watkinson, A., C. Cheng and C. Lim. “Oxygen-steam gasification of coals in a spouted bed” Canadian. Journal of
Chemical Engineers, vol.65, 1987, pp.791-798.

- 26 -
[84] Estephane J, Aouad S, Hany S, El Khoury B, Gennequin C, El Zakhem H, et al. “CO2 reforming of methane over Ni-
CO/ZSM5 catalysts” Aging and carbon deposition study. Int. J. Hydrogen Energ., vol.40, 2015, pp.9201-8.
[85] T. Hanaoka, T. Miyazawa, M. Nurunnabi, S. Hirata, K. Sakanishi “Liquid Fuel Production from woody biomass via
oxygen-enriched air/CO2 gasification on a bench scale” J. Jpn. Inst. Energy, vol.90, 2011, pp.1071-1080.
[86] Baratieri, M., Baggio, P., Fiori, L., Grigiante, M. “Biomass as an energy source: Thermodynamic constraints on the
performance of the conversion process” Bio resource Technology, vol.99, 2008, pp.7063-7073.
[87] Anjireddy Bhavanam and R. C. Sastry “Biomass Gasification Processes in Downdraft Fixed Bed Reactors: A
Review” International Journal of Chemical Engineering and Applications, vol.2, 2011, pp.6.
[88] Tae-Young Mun, Pyeong-Gi Seon, Joo-Sik Kim “Production of a producer gas from woody waste via air gasification
using activated carbon and a two-stage gasifier and characterization of tar” Fuel, vol.89, 2010, pp.3226-3234.
[89] Miccio F, Piriou B, Ruoppolo G, Chirone R. “Biomass gasification in a catalytic fluidized reactor with beds of
different materials” Chem. Eng J., vol.154, 2009, pp. 369-74.
[90] J.M. Seiler, C. Hohwiller, J. Imbach, J.F. Luciani, “Technical and economical evaluation of enhanced biomass to
liquid fuel process” Energy, vol.35, 2010, pp.3587-3592.
[91] Heidi C Butterman and Marco J Castaldi. “CO2 as a carbon neutral fuel source via enhanced biomass gasification”
Environmental science & technology, Vol.43, 2009, pp.9030-7.
[92] M. Pohorel´y, M. Jeremi´aˇs, K. Svoboda, P. Kamen´ıkov´a, S. Skoblia, and Z. “Beno. CO 2 as moderator for biomass
gasification” Fuel, vol.117, 2014, pp.198-205.
[93] Weil, S.A., S.P. Nandi, D.V. Punwani and J.L. Johnson. “Peat hydrogasification” Presented at 176th National Meeting
of ACS, Maiami, 1978.
[94] Li K, Zhang R, Bi J “Experimental study on syngas production by co-gasification of coal and biomass in a fluidized
bed” Int. Hydrogen Energy, vol.35, 2009, pp.2722-6.
[95] Eilhann E Kwon, Haakrho Yi, and Marco J Castaldi. “Utilizing Carbon Dioxide as a Reaction Medium to Mitigate
Production of Polycyclic Aromatic Hydrocarbons from the Thermal Decomposition of Styrene Butadiene Rubber”
Environmental science & technology, vol.46, 2012, pp.10752-10757.
[96] Eilhann E Kwon, Eui-Chan Jeon, Marco J Castaldi, and Young Jae Jeon. “Effect of carbon dioxide on the thermal
degradation of lignocellulosic biomass” Environmental science & technology, vol.47, 2013, pp.10541-7.
[97] M. Pohorel´y, M. Jeremi´aˇs, K. Svoboda, P. Kamen´ıkov´a, S. Skoblia, and Z. Beno. “CO2 as moderator for biomass
gasification” Fuel, vol.117, 2014, pp.198-205.
[98] Radmanesh, R., Chaouki, J., & Guy, C. “Biomass Gasification in a Bubbling Fluidized Reactor: Experiments and
Modeling” American Institute of Chemical Engineers Journal, vol.52, 2006, pp.4258-4272.
[99] Petersen, I., & Werther, J. “Experimental Investigation and modeling of gasification of sewage sludge in the
circulating fluidized bed” Chemical Engineering and Processing, vol.44, 2005, pp.717-736.
[100] de Andrés, J., Narros, A., & Rodríguez, M. “Air-steam gasification of sewage sludge in a bubbling bed reactor: Effect
of alumina as a primary catalyst” Fuel Processing Technology, vol.92, 2011, pp.433-440.
[101] Gómez-Barea, A., Leckner, B., Perales, A., Nilsson, S., & Cano, D. “Improving the performance of fluidized bed
biomass/waste gasifiers for distributed electricity: A new three-stage gasification system” Applied Thermal Energy,
vol.50, 2013, pp.1453-1462.
[102] Manyà, J.J., Aznar, M., Sánchez, J.L., Arauzo, J., & Murilla, M.B. “Further Experiments on Sewage Sludge Air
Gasification: Influence of the Nonstationary Period on the Overall Results” Industrial & Engineering Chemistry,
vol.45, 2006, pp.7313-7320.
[103] Mansaray K, Ghaly A, Al-Taweel A, Hamdullahpur F, Ugursal V. “Air gasification of rice husk in a dual distributor
type fluidized bed gasifier” Biomass Bioenergy, vol.17, 1999, pp.315-32.
[104] Qin Y, Feng J, Li W. “Formation of tar and its characterization during air-steam gasification of sawdust in a fluidized
bed reactor”. Fuel, vol.89, 2009, pp.1344-7.
[105] Skoulou V, Koufodimos G, Samaras Z, Zabaniotou A. “Low temperature gasification of olive kernels in a 5-kW
fluidized bed reactor for H2-rich producer gas” Int. J. Hydrogen Energy, vol.33, 2008, pp.6515-24.
[106] Gil J, Aznar M, Caballero M, Frances E, Corellas J. “Biomass gasification in fluidized bed at pilot scale with steam-
oxygen mixtures. Product distribution for very different operating conditions” Energy Fuels, vol.11, 1997, pp.1109-
18.
[107] Huiyan Zhang, Rui Xiao, Denghui Wang, Guangying He, Shanshan Shao, Jubing Zhang, and Zhaoping Zhong.
“Biomass fast pyrolysis in a fluidized bed reactor under N 2, CO2, CO, CH4 and H2 atmospheres” Bioresource
technology, vol.102, 2011, pp. 4258-64.

- 27 -
[108] Asadullah M, Miyazawa T, Ito S, Kunimori K, Koyama S, Tomishige K. “A comparison of Rh/CeO 2/SiO2 catalysts
with steam reforming catalysts, dolomite and inert materials as bed materials in low throughput fluidized bed
gasification systems” Biomass Bioenergy, vol.26, 2004, pp.269-79.
[109] Miccio F, Piriou B, Ruoppolo G, Chirone R. “Biomass gasification in a catalytic fluidized reactor with beds of
different materials” Chem. Eng J., vol.154, 2009, pp. 369-74.
[110] Basu P (2010) , Chapter 5 - gasification theory and modeling of gasifiers, in Biomass Gasification and Pyrolysis
Anonymous Boston: Academic Press, , pp. 117-165.
[111] Puig-Arnavat M,. Bruno J. C , Coronas A. (2010)., Review and analysis of biomass gasification models,
Renewable and Sustainable Energy Reviews, 12, pp. 2841-2851
[112] Melgar A., Pérez J. F., Laget H.,Horillo A. (2007) , Thermochemical equilibrium modelling of a gasifying process,
Energy Conversion and Management, 1, pp. 59-67.
[113] Jarungthammachote S. Dutta A. (2007), Thermodynamic equilibrium model and second law analysis of a
downdraft waste gasifier, Energy, 9, pp. 1660-1669.
[114] Zainal Z. A., Ali R., Lean C. H., Seetharamu K. N. (2001) Prediction of performance of a downdraft modeling for
different biomass materials, Energy conversion and Management, 42 , pp.1499-1515.
[115] Li X., Grace J. R., Watkinson A. P., Lim C. J. Ergüdenler A. (2001), Equilibrium modeling of gasification: a free
energy minimization approach and its application to a circulating fluidized bed coal gasifier, Fuel, 1 , pp. 195-207
[116] Koroneous C., Lykidou S. (2011), Equilibrium modeling for a downdraft biomass gasifier for cotton stalks biomass
in comparison with experimental data, Journal of Chemical Engineering and Materials Science, 2, pp. 61-68.
[117] Ratnadhariya J. K., Channiwala S. A. (2009), Three zone equilibrium and kinetic free modeling of biomass gasifier
– a novel approach, Renewable Energy, 4 , pp. 1050-1058.
[118] Bridgwater A. V. (2012), Review of fast pyrolysis of biomass and product upgrading, Biomass Bioenergy, 3, pp.
68-94.
[119] Centeno F., Mahkamov K., Silva Lora E. E., Andrade R. V(2012), Theoretical and experimental investigations of
a downdraft biomass gasifier-spark ignition engine power system, Renewable Energy, 1, pp. 97-108
[120] Koukkari P, Pajarre R. (2006), Introducing mechanistic kinetics to the Lagrangian Gibbs energy calculation,
Comput. Chem. Eng., 30, pp. 1189-1196.
[121] Jarungthammachote S., Dutta A. (2008), Equilibrium modeling of gasification: Gibbs free energy minimization
approach and its application to spouted bed and spout-fluid bed gasifiers, Energy Conversion and Management, 6,
pp. 1345-1356.
[122] Antonopoulos I., Karagiannidis, A. Gkouletsos A., Perkoulidis G.(2012), Modelling of a downdraft gasifier fed
by agricultural residues, Waste Manage., 4, pp. 710-718.
[123] Srinivas T., Reddy B.V., Gupta A. V. S. S. K. S.(2012) Thermal Performance Prediction of a Biomass Based
Integrated Gasification Combined Cycle Plant - Journal of Energy Resources Technology-ASME Transactions 134,
pp.1-9
[124] Mathieu P., Dubuisson R. (2002) Performance analysis of a biomass gasifier, Energy Conversion and Management,
43, pp. 1291-1299.
[125] Mitta N.,R.,Ferrer-Nadal S.,Lazovic A.,M., Perales J.F., Velo E.,Puigjaner L. (2006) Modelling and simulation
of a tyre gasification plant for synthesis gas production, Proceedings of 1 6th European Symposium on Computed
Aided Process Engineering and 9th International Symposium on Process Systems Engineering, Garmisch-
Partenkirchen Germany, pp. 1771-76.
[126] NikooM. B., Mahinpey N. (2008) Simulation of Biomass gasification in fluidized bed reactor using ASPEN PLUS,
Biomass and Bioenergy, 32, pp. 1245-1254.
[127] Doherty W., Reynolds A., Kennedy D. (2008) Simulation of a Circulating Fluidised Bed Biomass Gasifier using
ASPEN Plus a Performance Analysis, Proc. 21st International Conference on Efficiency, Cost, Optimization,
Simulation and Environmental Impact of Energy Systems, Krakow, Poland. Pp.1241-1248.
[128] Fr´ed´eric Paviet Florent Chazarenc Mohand Tazerout (2009) Thermo Chemical Equilibrium Modelling of a
Biomass Gasifying Process Using ASPEN PLUS, International Journal Of Chemical Reactor Engineering, 7, Article
A 40, pp.1-16.
[129] Hannula I.,Kurkela E. (2010) A semi-empirical model for pressurised air-blown fluidised-bed gasification of
biomass, Bioresource Technology, 101, pp. 4608-15.
[130] Atnaw S. M., Sulaiman S. A., and yusup S. (2011) A Simulation Study of Downdraft Gasification of Oil-palm
Fronds using ASPEN PLUS , Journal of Applied science 11(11), pp 1913-1920.

- 28 -
[131] Ramzan Naveed,Asma Ashraf, Shahid Naveed,Malik Abdullah (2011) Simulation of hybrid biomass gasification
using ASPEN PLUS, A comparative performance analysis for food, municipal solid and poultry waste, Biomass and
Bioenergy,35(9) pp.3962-3969.
[132] Damartzis T., Michailos S., Zabaniotou A. (2012) Energetic assessment of a combined heat and power integrated
biomass gasification–internal combustion engine system by using Aspen Plus , Fuel Processing Technology, 95, pp.
37-44.
[133] Chen C., Jin Y. Q., YanJ. H., Chi Y. (2013) Simulation of municipal solid waste gasification in two different types
of fixed bed reactors, Fuel, 103, pp. 58-63.
[134] Pavlas M., Stehlík P., Oral J., Klemeš J., Kim J. K., and Firth B. (2010) Heat Integrated Heat Pumping for
Biomass Gasification Processing, Applied Thermal Engineering , 30, No. 1, pp. 30-35.
[135] Sipöcz Nikolett, Finn Andrew Tobiesen, Assadi Mohsen (2011), Applied Energy, 88, 2368-2376.
[136] Maria Puig-Arnavat , Alfredo J. Herna´ndez , Carles Bruno Joan, Alberto Coronas (2013) Artificial neural
network models for biomass gasification in fluidized bed gasifiers , Biomass and Bioenergy, 49, pp. 279-289.

Sources for the Figures

Figure 1.0
Ptasinski, K. J. (2015). Efficiency of biomass energy: An exergy approach. John Wiley.

Figure 1.2:
Basu, P. (2009). Combustion and Gasification in Fluidized Beds. New York.

Figure 1.3:
Li, X., Grace, J. R., Lim, C. J., Watkinson, A. P., Chen, H. P., and Kim, J. R.,(2004) Biomass
gasification in a circulating fluidized bed, Biomass Bioenergy, 26, 171–193.

Figure 1.4:
Kaiser, S., Loffler, G., Bosch, K., and Hofbauer, H., (2003). Hydrodynamics of a dual fluidized
bed gasifier, Part II: simulation of solid circulation rate, pressure loop and stability, Chem. Eng.
Sci., 58, 4215–4223.

Figure 1.5:
Source: http://www.hauserman-engineering.com/Gasification.html

- 29 -

Вам также может понравиться