Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

ch5-ed

Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

<RRH>Modeling of Gasifiers for Biohydrogen Production

<CN>CHAPTER 5

<CT>Modeling of Gasifiers for Biohydrogen Production

<AU>Abinash Mahapatro1, Binayak Pattanayak1, Alok Kumar2, and Krushna

Prasad Shadangi3

<AFN>1Department of Mechanical Engineering, SOA Deemed to be University

Bhubaneswar, Bhubaneswar, India

<AFN>2Department of Mechanical Engineering, Indian Institute of Technology Bombay,

Powai, Mumbai, India

<AFN>3Department of Chemical Engineering, Veer Surendra Sai University of

Technology, Burla, Sambalpur, Odisha, India

<ABS>ABSTRACT

In this current era, energy is a prime requirement for mankind. The primary source

of energy across the globe is mitigated from coal through the combustion route. The major

drawback of this route lies in the emission characteristics. To reduce the emission and

meet the current energy demand, it can be produced through a cleaner way through the

gasification route. In the gasification process, syngas is obtained, which mainly consists of

hydrogen, methane, carbon monoxide, and higher order of hydrocarbons. Among these,

hydrogen is considered a cleaner fuel that has the potential to meet the energy demand.

Satisfactory results are not available in the computational modeling of hydrogen

production in the open literature. The research methodology of hydrogen production

through the thermochemical process is outlined. The current chapter is an effort to

represent the flow path of hydrogen production through experimental and modeling routes.

<KW>Keywords: gasification, hydrogen, gasifier, modeling, syngas


<H1>5.1 INTRODUCTION

For the production of power, coal is considered the primary energy source

throughout the world. Particularly for Asian countries like India and China coal is

abundantly used for power production and generation. About 72% of electricity in China

and 58% in India is produced from coal (Key World Energy Statistics, 2018). In

developed countries like Japan and the USA also about 32% and 40% of electrical power

are generated from coal respectively (Key World Energy Statistics, 2018). Apart from coal

different biomasses like a corn husk, rice husk, rice straw, and sawdust are abundantly

available in India and can be used as a potential source for power production (Mallick et

al., 2020). The use of biomass for power production has advantages like reduced

dependency on fossil fuels like coal and reduced environmental pollution. The conversion

of fossil fuel to the generation of power takes place through the combustion and

gasification route. The combustion of coal leads to the emission of other harmful gases

like CO2, NOX, SOX, and particulate matter which creates an environmental concern. The

gasification route refers to the processes in which syngas is generated which are used for

the production of power using the thermochemical process and biological track using a gas

engine or fuel cell, respectively. Gasification occurs at relatively moderate temperatures

reducing the exergy losses to improve the energy efficiency of the system. The product of

gasification is syngas consisting of CO, H2, N2, and CH4, having low to medium calorific

value. Moreover, water consumption in the gasification process is less than that of

combustion. The attractive feature of the gasification is that air pollution is drastically

reduced along with a reduction of particulate emission. Thus, the development of new-

generation gasifiers to handle multi-feed fuel got attention in recent years. Biohydrogen is

one of the clean renewable energy resources in which hydrogen is produced and that can

be used for the generation of electricity using the fuel cell system. It has the potential to

replace the fossil fuel and mitigate the energy demand in next few decades.
The research activity in the gasification process consists of both the experimental

and numerical approaches. The experimental approach requires an experimental setup, i.e.,

the gasifier and the components viz. compressor, flow meter, distributor plate, riser,

heating unit, flow control valves, temperature and pressure-measuring devices, etc. The

use of the above components in the gasifier makes the system complex and higher

installation cost. Apart from this, the material property also puts a limitation on the

temperature of gasification. The pressure inside the gasifier is limited by the size of the

compressor while the temperature is also limited by safety issues in the laboratory scale

units. In this context, due to such limitations, the researchers have focused on the

numerical investigation of the gasifier.

The advantage of numerical investigation over the experimental work is that there is

no limitation on pressure or temperature limits and scale-up of the unit may be done from

laboratory to industrial-scale units. The major challenges in the numerical approach are the

selection of accurate flow models, governing equations, boundary conditions, and solution

controls. The study of numerical approach is conducted in the commercial software

packages like StarCCM+, ANSYS, COMSOL, and Aspen+. These software packages are

very user-friendly and each has certain advantages over the other.

<H1>5.2 GASIFICATION MECHANISM

Gasification is the transformation of carbonaceous fuel, called inventory or

feedstock, to the gaseous product with useful heating significance. This does not include

the combustion, because the product that forms the combustion is flue gas, not fuel gas so

the combustion product does not have a residual heating value. It incorporates the

advances of pyrolysis, partial oxidation, and hydrogenation. Commonplace inventories are

coal, wood, strong waste, and so on, which comprise carbon, hydrogen, and oxygen.

Generally, the medium of gasification is air, oxygen, steam, and CO2. The producer gas

obtained from air gasification has a low heating value and it is around 3.5–6 MJ/m3.
However, the producer gas obtained from pure oxygen gasification has a high heating

value and it is around 10–18 MJ/m3, yet it is disadvantageous because the cost of pure

oxygen production is high (Frigo and Spazzafumo, 2018).

The process of gasification can be classified into four different groups or four

regions: drying region, pyrolysis region, oxidation region, and hydrogenation. Though in

certain kinds of gasifiers, for example, spouted bed and fluidized bed, these regions cannot

be isolated because of excellent mixing of feed material, and every one of the four cycles

can be viewed as happening at the same time all around the reactor (Shayan et al., 2018).

The concentration of produced gas is a function of the temperature and pressure, and the

temperature within the gasifier mainly depends on the reactions of gasification. These are

the main reaction that occurs inside the gasifier.

Gasification reaction includes both the exothermic and endothermic reactions.

Typically, the process of gasification that utilizes air or oxygen as an oxidizing medium is

self-providing the heat, which is essentially consumed in the drying of feedstock,

pyrolysis, and hydrogenation. From the given reaction if carbon present in the feedstock

can be completely converted, the major yields from the cycle would be CO, CO2, H2, CH4,

and H2O. N2 is additionally a critical segment for the situation of air gasification.

The portable downdraft gasifier can be the best solution for the establishment and

focus among the public. The gasifier may be taken to diverse cotton gins and agricultural

enterprises for the generation of heat and power (Capareda et al., 2007). The content of

moisture in the inventory is directly associated with its consumption. If the content of

moisture in the inventory enhances, then the rate of consumption as well as the

temperature inside the gasifier decreases to a certain limit. However, an increase in the

flow rate of air enhances the consumption of inventory as well as the temperature of the

gasifier (Sheth and Babu, 2009). A superficial velocity also affects the gasifier
performance, if it is lower then the rate of pyrolysis decreases, and production of tar boosts

within the gasifier, whereas a high superficial velocity of air enhances the pyrolysis

reaction. Due to the increase in superficial velocity of air a decrement in the

manufacturing charge of the coal in addition to tar will enhance the flow rate of gas (Jong

and Hein, 2005). The temperature enhances within the gasifier with the boom in gas flow

rate surge in the ratio of air and biomass. The enhanced temperature in the gasifier

consequences the better conversion of CO2 and H2O into CO and H2 which enriches the

calorific value of producer gas as well as the efficiency of product gas (Sharma, 2011).

The material of biomass (shell) like coconut shell, groundnut shell, and rice husk can be

transformed into combustible gas at about 800 °C with the technique of gasification. High-

quality producer gas is obtained for coconut shell which has the calorific value 23 % more

than that of a ground nutshell as well as 45 % more than rice husk (Mahmoudi et al.,

2011). The main problem with the downdraft gasifier is its tar content, 48 g/Nm3 average

tar is produced for the duration of the gasification (Dogru et al., 2002).

<H1>5.3 GASIFICATION USING FLUIDIZED-BED SYSTEM

A fluidized-bed gasifier (FBG) is one of the promising devices for sustainable

gasification of solid feed materials as long as the same can be fluidized with sub-

stoichiometric air. Low-grade coal, rice husk, corn husk, sawdust, sugarcane bagasse, etc.

can be handled by this gasifier, significantly reducing the cost of grinding and ball milling

of the same. More than 700 atmospheric circulating fluidized bed (ACFB) boilers are

installed in the world for power generation from coal. In contrast, pressurized circulating

fluidized bed (PCFB) is gaining importance since the late 1990s due to its compactness

and high power density. However, only a few studies on the gasification of low-grade coal

and biomass are reported in the literature using PCFB. Hence there is a call for a thorough

investigation on the gasification of low-grade coal and loose biomass in PCFB.


Fluidized-bed gasification is one of the promising technologies for the conversion of

carbonaceous content in feedstock into syngas. This process eventually occurs inside the

reactor of the fluidized bed, where hot air is directly mixed with the feed materials such as

coal and different types of biomasses resulting in a series of chemical reactions to produce

syngas (Basu and Mettanant, 2009). The syngas typically comprises different flammable

gases such as CO, CH4, H2, and CO2. To analyze the composition in a more qualitative

way and to couple these with the effectiveness of the experimental facility, usually,

various parameters such as dry gas yield (Y), lower heating value (LHV), cold gas

efficiency (CGE), and carbon conversion efficiency (CCE) are estimated (Mallick et al.,

2019). The significant input parameters associated with the operation of gasifiers are

equivalence ratio, operating pressure, temperature, feed material, and steam-to-fuel ratio

(Pinto et al., 2009; Seo et al., 2010; Wu and Williams, 2010). The ash formed during the

gasification process is also one of the important parameters. It is, in fact, the by-product of

this process that can further be used as a marketable solid. The gasification process may be

enhanced by using different types of catalysts such as dolomite which absorbs the oxides

formed from sulfur. In this section, a brief discussion of available literature incorporating

the important input as well as output parameters concerning gasification studies is

presented.

Padban et al. (2000) investigated the tar formation of different sawdust in PBFB at a

pressure of 12 bar. They used the recalcined magnesite as bed material and observed that

an increment in pressure gives rise to a reduction in devolatilization, thereby decreasing

the tar formation. Mayerhofer et al. (2012) also studied the various parameters like

pressure, temperature, and steam on tar production which affect the fluidized-bed

gasification. Olivine having an average particle size of 0.25 mm was used as bed material.

They have carried out experiments at an elevated pressure of 2.5 bar and two different

temperatures 750°C and 790°C. The result reflected an increase in CO and H2


concentration along with a decrease in CO2 and CH4 concentration with the rise in the

temperature. Abdoulmoumine et al. (2014) investigated the gasification characteristics of

pine wood sawdust in the BFB gasifier. The effect of temperature and ER on the

production of synthesis gas along with the concentration of tar in the syngas was

investigated. They observed that with an enhancement of temperature, gas yield increases

while the concentration of tar decreases in the syngas. Berrueco et al. (2014) also analyzed

the formation of tar during the gasification of biomass in the PBFB gasifier at an elevated

pressure of 5 bar by considering the effect of dolomite for two different ranges of

temperature 750°C and 850°C. They used sand and dolomite sieved of particle size 150–

200 µm as the bed material. The use of dolomite in place of sand helps the concentration

of CO and CO2 to increase with an increase in temperature. By increasing the operating

pressure up to 10 bar, Berrueco et al. (2014) studied the gasification characteristics and tar

formation of torrefied woody biomass in a lab-scale PBFB reactor at an operating

temperature of 850°C. They observed that with an increase in pressure, the tar formation

increases. Sahoo and Ram (2015) investigated the biomass characteristics in the BFB

gasifier for sugarcane bagasse and used different parameters to calculate the effect on

syngas production. They had used dolomite, sand, and red mud as a bed material and

found that when the mixture of sand and red mud was used in the ratio of 1:1, H2 yield

was maximum as compared to other contemporary bed materials. They have also observed

that with an increase in temperature from 500°C to 850°C, the concentration of H2 yield

enhances, whereas the concentration of CO, CO2, and CH4 reduces. Zeng et al. (2016) also

investigated the gasification of pine sawdust in a dual fluidized bed having two separate

reactors (fuel reactor and steam reactor) for gasification of sawdust. They found that in a

fuel reactor with an increase in temperature from 700–800°C composition of H2, CO, and

CH4 increases while CO2 composition in syngas decreases. The CGE was found to be

maximum viz. 77.2% for riser temperature of 820°C. However, in the steam reactor with
an increase in temperature from 700–970°C, the concentration of H2 increases while CO,

CO2, and CH4 decrease. CGE in the steam reactor was reported to have a decreasing trend

from 81 to 77% with an increase in temperature. Nevertheless, in the near past, many

researchers have been involved with analyzing the gasification characteristics in BFB and

PBFB reactors by considering different biomasses, bed materials, and operating

parameters (Kitzler et al., 2011; Han et al., 2013; Figueroa et al., 2014; Suksuwan et al.,

2018). However, similar studies related to BFB or PBFB along with coal as inventory

have not been explored too much extent.

Huang et al. (2003) analyzed the characteristic of Chinese Fugu Coal with an air-stream

mixture in BFB at a temperature of 880–980°C and a pressure range from 5 to 14 bar.

Xiao et al. (2006) focused on the high-temperature gasification of coal in the pressurized

spout gasifier. During the gasification, they varied different parameters like pressure (1–5

bar), gasifying agent temperature (300–700°C), ER (0.28–0.48), and steam-to-coal ratio

(1.1–3.8 kg/h). They have reported that with an increase in pressure, the concentration of

H2 and CO increases while CO2 and CH4 remain almost constant. Analysis of HHV, cold

gas efficiency, and carbon conversion efficiency reveals an increment of values from

pressure 1 to 3 bar. A similar trend was also perceived by Qin et al. (2007). They

examined the gasification of coal in both the BFB and PFB reactors. They noticed that the

conversion of coal into syngas increases with temperature and residence time for both

BFB and PFB reactors. The gasification characteristics of underground coal in the BFB

gasifier were studied by Konstantinou and Marsh (2015). They surveyed the effect of

reactant gas pressure on the syngas composition. They found that the concentration of H2

is maximum at 9 bar and 16.5 bar pressure respectively. Sánchez et al. (2016) also studied

the gasification characteristics of coal and developed a model for coal gasification in the

PFB reactor having pressure range 1–6 bar and temperature range 920–1075°C. Gul et al.

(2018) described the PBFB reactor gasification for Turkish lignite coal where the
operating temperature of the gasifier was varied between 870 and 880°C and pressure

from 2.4 to 2.7 bar. They have used silica sand as a bed material and a mixture of oxygen,

steam, CO2, and air as a gasifying agent.

Unlike the comprehensive investigation of PBFB gasifiers, PCFB has been less

explored by researchers. In the operating pressure range of 100–160 kPa, Li et al. (2019)

used coal as bed inventory to study the equilibrium modeling of the PCFB gasifiers. The

authors reported that increased operating pressure results in increased CO2 and H2

concentration whereas CO content is found to decrease. The decrement in CO content

along with increment in concentration of hydrocarbon and CO2 has been reported with an

increase in pressure. A 300 kW CFB pilot plant was studied by García-Ibañez et al. (2004)

using silica sand and olive oil as bed material and feed material respectively. They

reported that the concentration of H2, CO, and CH4 decreased with an increase in ER while

the concentration of CO2 is enhanced due to shifting of gasification to combustion at

higher ER. Gas yield and CCE showed an increasing trend whereas LHV decreases with

an increase in ER. A PCFB gasifier running on a combined power cycle was studied by

Srinivas et al. (2006) using coal as a fuel. Gasifier temperature was reported to increase

from 920°C to 980°C with a rise in the pressure from 400 to 3200 kPa. Similarly, the rise

in gasifier pressure results in a decrease in CO2 mole fraction and irreversibility of the

gasifier. Srinivas et al. (2009) investigated PCFB riser for thermodynamic equilibrium

modeling of biomass gasification. They have considered different biomass like rice husk,

sawdust, solid waste, and manure and thereby studied the effect of air-fuel ratio (AFR),

steam-to-fuel ratio (SFR), and gasifier pressure on the syngas production. In the operating

pressure range of 100–500 kPa, the gasification of bituminous coal was conducted by

Duan et al. (2010) in a lab-scale PCFB riser. The authors reported an increase in oxygen

flow rate with an increase in operating pressure. The increase in oxygen production results

in an enhanced gasification rate and increased temperature up to an ER of 0.39. However,


for subsequent ER, combustion plays a significant role. The LHV of syngas was observed

to have a maximum value for the stoichiometric ratio of 0.35. The authors observed a

direct relationship of temperature with gasification efficiency. They have also identified

that the syngas concentration of H2 and CO increases, whereas N2 and CO2 decrease when

the temperature rises. Guan et al. (2007) investigated the impact of reaction pressure based

on the thermodynamic equilibrium analysis of biomass. They observed that the fraction of

syngas composition remains constant under the pressure range 0.1 MPa–0.6 MPa. Above

0.6 MPa the concentration of H2 increased rapidly and achieved the maximum value at a

pressure of 1.3 MPa, and thereafter the concentration of H2 starts decreasing slowly with

an increase in pressure because at elevated pressure methanation reaction takes place.

With the increase in operating pressure, the CH4 concentration increases continuously

whereas CO and CO2 production decreases. Long et al. (2013) examined the production of

H2-rich syngas through PBFB gasification of sawdust and used silica sand as bed material.

The authors reported that with an increase in operating pressure, the CO and CO2

production decreases while that of H2 and CH4 increases. The reason for the above

observation is due to a shift in water gas reaction in enhanced pressure conditions.

Sugarcane bagasse was gasified by Figueroa et al. (2014) in a BFB gasifier. They reported

that the synthesis gas production increases in enhanced ER condition due to increase in

temperature. For an ER of 0.28, about 95 wt% of synthesis gas was produced. Kitzler et al.

(2011) experimented on wood pellet gasification with the use of air, steam, oxygen, and

CO2 as gasification agents and olivine as bed material. For their experimental observation,

they have varied different parameters like temperature, pressure, and air ratio. The authors

observed that operating temperature above 1093 K and pressure below 400 kPa are

suitable for higher H2 concentration in synthesis gas production. The authors also

mentioned that CO concentration decreases with increase in pressure, while CO2 and CH4

are observed to follow the opposite trend. They also noticed that the concentration of CH4
and CO2 increases with pressure, while the concentration of CO follows the opposite

trends to CO2. In the operating pressure range of 100–2600 kPa, Li et al. (2018)

experimented on the co-gasification of meat bone meal and coal. The authors reported that

increase of pressure produces favorable condition for the enhancement of gasification rate

and fuel production while decreasing slag. They also found that the carbon conversion

efficiency (CCE) improved with an increase in pressure. Seo et al. (2010) investigated the

gasification and co-gasification of Indonesian Tinto sub-bituminous coal and sawdust in a

dual circulating fluidized bed reactor. They observed that at 900°C, temperature gas yield

increases with an increase in biomass ratio. They also noticed that the concentrations of H2

and CO were maximum for a biomass ratio of 0.5 at 900°C. CGE and calorific value was

found to be maximum for 50% blend ratio which was higher than individual gasification

of biomass and coal.

In conjunction with the experimental assessment, some of the earlier research works

have also been diverted toward numerically simulating these gasifiers. Investigation of

gasification characteristics of sugarcane bagasse was studied numerically by Pellegrini and

de Oliveira, (2007). Based on the exergy analysis, a numerical study was developed. They

reported that an increase in ER concentration of CO decreases, and CO2 increases because

higher ER gasification converts in the combustion reaction. At an ER of 0.25, the

concentration of H2 was found to be maximum and it reaches up to 22%, and at the same

point, LHV and CGE were also found to be maximum. They have also noticed that an

increase in moisture content results in decreased CO concentration while CO2 and H2

follow the opposite trend. On the other hand, the LHV of syngas was found to be

decreasing. Li et al. (2009) developed a numerical model for coal gasification of

pressurized spout-fluid bed and simulations were done at 1, 3, and 5 bar pressure and a

steam temperature of 350°C. Camacho Ardila et al. (2012) numerically examined the

model of gasification of sugarcane bagasse in the CFB reactor using Aspen Plus. Zogola et
al. (2014) modeled a coal gasifier and tried to simulate the effect of various factors like

pressure, temperature, and gasifying agent on the composition of the syngas. They have

considered temperature ranging from 500–1500°C and pressure from 1–35 bar for their

simulation and used steam and oxygen as a gasifying agent. They have reported that with

the rise of operating pressure conditions, the mole fraction of H2, CO, LHV, and CGE

decreases while the mole fraction of CH4 and CO2 increases. However, with the rise in

temperature, the LHV, CGE, and the mole fraction of H2 and CO, increase whereas the

mole fraction of CO2 decreases. Adeyemi et al. (2017) compared the experimental results

of gasification analysis for Kentucky coal and woody biomass with numerical simulation.

The numerical simulation was carried out using ANSYS FLUENT for 2D and 3D

geometry, and the results were compared. The Eulerian–Lagrangian model was used as the

viscous model. The effect of fuel type, pressure, temperature, and ER on syngas

composition was studied. The numerical results were found to be accorded with the

experimental results. Karmakar and Datta, (2011) numerically studied the gasification of

rice husk in BFB and compared the results with the experimental observation. They

noticed that the concentration of H2 and CO is enhanced, while CO2 and CH4 decrease

with an enhancement of temperature. On the other hand, it was found that with an increase

in steam-to-biomass ratio, the concentration of CO2 and H2 improved while CO and CH4

concentration reduces. Both temperature and steam-to-biomass ratio favored the CGE and

CCE. Klimanek and Bigda (2018) modeled a PCFB gasifier in Barracuda and a fluent

simulation tool for coal gasification in an operating pressure range of 100–1500 kPa. The

results obtained from the simulation are compared with those of a small-scale PCFB coal

gasifier. Motta et al. (2019) analyzed the gasification of sugarcane bagasse in Aspen

PlusTM and used various operating parameters like temperature (750–950°C), pressure (1–

15 bar), steam-to-biomass ratio (0.5–1.5), and moisture content (10–30%) to obtain syngas
composition. Table 5.1 highlights the summary of the literature related to gasification

studies by various authors.

TABLE 5.1 Gasifier Specification.

Investigators Gasifier Operating conditions Optimum results

specification

Jong et al. BFB Air The concentration of


(1999)
ID: 0.4 m Coal and Biomass, Sand CH4 and C2H2 is a

H: 4 m P: 1000 kPa, T: 973–1173 K function of the air/fuel

ratio.

Sjöström et al. BFB Oxygen Char production

(1999) ID: 144 mm Birch wood: 1–3 mm decreases while gas

H: 600 mm Bituminous coal: 1–1.35 mm production increases

Silver sand and olivine sand with biomass ratio.

P: 400 kPa, T: 973–1173 K

Padban et al. BFB Air Increasing the pressure

(2000) ID:102 mm Woody biomass reduced the

H: 3.3 m Precalcined magnesite devolatilization and

dp: 0.180 mm decreased tar

P: 500–2000 kPa, T: 1123– production.

1223 K

Li et al. (2001) CFB Air Hydrocarbon

Modeling Coal production increases

ID: 100 mm dp< 0.6 cm with an increase in

H: 6.3 m P :100–160 kPa, T:1023–1173 pressure.

K
Huang et al. BFB Air–steam H2: 14.57–18.08%
(2003) ID: 200 mm Chinese Fugu Coal CH4: 2.91–2.55%

dp: 0.086–0.3 cm

P: 530-1400 kPa; T: 1153–

1253 K

VS: 80–100 cm/s

Kurkela et al., CFB Air, Steam, CO2

ID: 154 mm Different biomass

H: 7.9 m Al2O3, Silica sand, Limestone,

Dolomite, MgO

P:100–130 kPa,T: 873–1273

Umf : 100–500 cm/s

Xiao et al. Spout fluid Air–steam H2: 10.6–15.2 (vol%)


(2006) bed Chinese Xuzhou coal CH4: 2.3–2.4 (vol%)

ID: 80 mm dp: 0.68×10–3 mm

H: 1.45 m Silica sand—dp: 0.75×10–3

mm

P: 100–260 kPa,T: 1143–1303

ER: 0.36–0.42

Guan et al. CFB Steam Above 6 bar H2


(2007) Biomass, corn stalk increases rapidly up to

P: 100–3000 kPa 13 bar and then

T: 923–1473 K decreases while CH4


continuously increases

with pressure.

Qin et al. BFB Air-steam With an increase in


(2007) ID: 34 mm Sawdust pressure, H2 increases

H: 500 mm dp: 80×10–3 –124×10–3 mm while tar content

P: 500–2000 kPa, T: 973– remains constant.

1173 K

Fermoso et al. BFB Steam–Oxygen The concentration of H2


(2009) ID: 13 mm Coal, biomass, petroleum increases with the

H: 305 mm coke increase in temperature.

P: 500–2000 kPa,T: 1123–

1273 K

O/C: 0.15–1.3

Li et al. (2010) Simulation of Steam–air The increase in


pressurized Chinese bituminous coal and pressure–volume
spout-fluid Xuzhou coal fraction of H2 and CH4
bed P: 100–500 kPa, T: 543–623 increases

A/C: 1.65–1.82

Srinivas et al. CFB Air–steam The concentration of H2


(2009) Rice husk, Sawdust, Solid and CH4 increases with

waste an increase in the S/F

P: 1000–2000 kPa, T: 1023– ratio.

1173 K
Valin et al. BFB N2, steam With an increase in
(2010) ID: 124 mm Sawdust pressure H2, CH4, and

H: 2.5 m dp: 10–20 cm dry gas yield increases

P: 200–400 kPa,T : 1073 K

Duan et al. CFB Air–steam H2: 10.55–13.62 vol%


(2010) ID: 70 mm Chinese bituminous coal

H: 6.8 m dp: 0.05 cm

P: 100–500 kPa, T: 673–973

Kitzler et al. PBFB Air–steam–Oxygen The concentration of H2


(2011) ID: 80 mm Olivine was found maximum

H: 500 mm dp: 500×10-3 mm for 3 bar at above

P: 100–600 kPa 850ºC. For CH4

T: 1023–1193 K production, higher

pressure is required

Mayerhofer et BFB Steam The concentration of H2


al. ( 2012)
ID:154 mm Wood pellets of 0.8 cm dia increases while CH4

H: 1500 mm Olivine—dp: 0.25 mm decreases with an

P: 100–250 kPa,T: 1023–1113 increase in temperature.

Han et al. BFB Cao, steam H2: 50.2–62.4%


(2013)
ID: 56 mm Sawdust—dp: (0.15–0.23)×10- CH4: 12.5–14.2%

H: 2300 mm
3 mm

Silica sand dp: 0.23–0.38 mm


P: 100–400 kPa, T: 803–1033

Berrueco et al. BFB Steam–Oxygen H2 yield decreases


(2014) ID: 23.8 mm Torrefied biomass while the content of

H: 404 mm (0.25–0.5 mm CH4 increases with an

P: 100–1000 kPa, T: 1123 K increase in pressure

Berrueco et al. BFB Air–steam H2, CH4 and dry


(2014) ID: 23.8 mm Torrefied biomass concentration of gas

H: 404 mm 0.250–0.550 mm increases with an

Sand and Dolomite 0.15–0.2 increase in temperature.

mm

P: 500 kPa, T: 1023–1123 K

Zogola et al. Simulation Air, Oxygen, steam H2 production


(2014) decreases while CH4
Coal

P: 100–3500 kPa,T : 773– increases with the

1773 K increase in pressure.

Sánchez et al. BFB Air–steam Effect of pr. variation


(2016) Simulation coal for H2 and CH4 was

P: 1000–5000 kPa, T: 923– very negligible

1073 K

Adeyemi et al. BFB Air Formation of H2


(2017) ID: 66 mm Kentucky coal increases with

H: 1540 mm woody waste temperature

dp: 0.100 mm enhancement.


P: 100–300 kPa; T: 923–1373

Gül et al. BFB Air, Oxygen, steam and CO2 H2: 31.5–34.2%
(2018) ID: 300 mm Coal—dp: 0.05–0.1 cm CH4: 4–4.6%

H: 3500 mm Silica sand—dp: 0.644 mm

P: 240–270 kPa,T: 1123–1203

Klimanek and CFB Air–steam The concentration of H2


Bigda ( 2018) modeling Coal is less in the analyzed

ID: 80 mm P: 100–1500 kPa, T: 1273 K system compared to the

conventional reactor.

Li et al. (2019) CFB


Oxygen–steam The rate of gasification

L: 300 mm JCA coal and MBM increases with an

W: 30 mm dp < 6000 μm increase in pressure.

H: 1.45 m Polystyrene—dp: 1.32 µm

P: 100–2600 kPa, T: 1173–

1673 K

The ongoing discussion trend of gas composition with operating parameters is given

in the tabular form (Table 5.2). From the table, it is observed that with an increase in

operating parameters such as equivalence ratio, temperature, streamflow ratio, and

pressure, it is difficult to distinguish the trends of the syngas. No particular trends of

syngas are observed and the trend is difficult for different operating conditions. The trends

are also a function of gasifier specifications. Hence, it is more important to investigate the

gasification study under a pressurized condition.

TABLE 5.2 Impact of Operational Parameters on Gas Composition.


Operating Gas composition
Investigators

parameters (↑) H2 CH4

ER ↓ ↑↓
Li et al. (2001)
T ↑ ↓

T ↑↓ ↑
Huang et al. (2003)
PR ↑ ↑↓

García-Ibañez et al. ER ↑↓ ↓

(2004)
T ↑ ↓

ER ↑↓ ↓

T ↑ ↑
Xiao et al. (2006)
PR ↑≈ ≈

Qin et al. ( 2007) T ↓↑ --

T ↑↓ ↓≈

PR ≈↑ ≈↑
Guan et al. (2007)
S/F ↑ ↓

Pellegrini and de ER ↑↓ ↓

Oliveira, (2007)
Moisture ↓ ↑

ER ↓ ↓

PR ↓ ↑
Srinivas et al. (2009)
S/F ↑ ↑

ER ↓ ↓
Li et al. (2010) PR ↑ ↑

ER ↑ ↓

T ↑ ↓

Duan et al. (2010)


PR ↑ ↑

S/F ↑ ↓

T ↑ ↑

Seo et al. (2010)


S/F ↑ ↓

ER ↓ ↓
Kitzler et al. (2011)
S/F ↑ ↑

Karmakar and Datta, T ↑ ↓

(2011)
S/F ↑ ↑

T ↑ ↓
Mayerhofer et al.
PR -- ↑
(2012)
S/F ↑ ↓

ER ↓ --

Ardila et al. (2012) T ↓ --

S/F ↑ --

T ↑ ↓

Han et al. (2013) PR ↑ ↑

S/F ↑ ↓

Żogała, (2014) ER ↑ --
T ↑≈ ↓

PR ↓ ↑

S/F ↑ ↑

Abdoulmoumine et al. ER ↓ ↓

(2014)
T ↑ ↓

Berrueco et al. (2014 a) T -- ↑

Berrueco et al. (2014 b) PR ↓ ↑

ER ↑ ↑
Figueroa et al. (2014)
T ↑ ↑

Konstantinou and
PR ↑↓ ↑
Marsh ( 2015)

Sahoo and Ram (2015) T ↑ --

ER ↓ ↓

T ↑ ↓
Sánchez et al. (2016)
PR ↓ ↑

S/F ↑ ↓

T ↑ ↑
Zeng et al. (2016)
S/F ↑ ↑

ER ↑ --

Adeyemi et al. (2017) T ↑ --

PR ↑ --
<H1>5.4 MODELING

In this section, the numerical work for the production of syngas in fluidized bed

gasifiers is discussed. Gasification of biomass is an appealing era for the transformation

of numerous kinds of biowastes to energy and it is an ecological process to produce

hydrogen (Barba et al., 2016). Biowastes gasification has been explored

in numerous studies of performance analysis. Though just a limited work has been done

on performance analysis for the production of hydrogen through gasification by Shayan et

al. (2018), Safarian et al. (2020) developed a model for downdraft air gasifier of waste

biomass using kinetic free equilibrium model to analyze the gasification performance of

garden waste, paper mixed waste, and timber and wood waste. The results indicated that at

lower temperatures (less than 500°C), the syngas has a low concentration of hydrogen due

to the remains of unburnt carbon and methane. At higher temperatures, oxidization of

carbon and reverse methanation yield higher concentrations of carbon monoxide and

hydrogen gas, respectively, in all three feedstocks. The study indicated that appropriate

conditions for producing the highest power efficiency are gasifier temperatures of 900–

1000°C for all waste and an equivalence ratio between 0.2 and 0.3, 0.4–0.5, and 0.35–0.45

for timber and wood, paper mixed, and garden wastes, respectively. The corresponding

power efficiencies from gasification are 45, 26, and 16% for timber and wood, paper

mixed, and garden wastes, respectively. For the comprehensive evaluation of hydrogen

through water–gas shift reactors, various modeling processes based on thermodynamic

equilibrium, kinetics, computational fluid dynamics, and synthetic neural networks

(ANNs) may be evolved.

The fashions derived via equilibrium strategies are impartial of the gasifier shape,

so maybe implemented for ideal structures and typical thermodynamic characteristics. But,

for a broadly complex method, correct kinetic parameters are required that are used

in kinetic modeling. In calculations depending on CFD, various energy equations, mass,


momentum, and species via a precise region of the gasifier are solved concurrently and

then only it gives correct temperature as well as concentration. The methods primarily

based on ANNs require a large number of records after which use a fixed of mathematical Commented [CE1]: AU: Please check the sentence ‘The
methods primarily based on…’ for clarity and amend this if required.

regressions for correlations between input and output information by Safarian et al.

(2020).

This technique has currently received a hobby due to the fact it can estimate nonlinear

functions without the requirement of the mathematical description of phenomena over

the gadget. Consequently, ANN models are appealing for final results prediction, while

critical interactions of complex nonlinearities are in the facts set, along with the

conversion of biomass (Capizzi et al., 2020).

One of the pleasant methods to simulate the overall performance of fluidized-bed

biomass gasification is the quasi-equilibrium technique (Puig-Arnavat et al., 2013), which

presents a greater accurate prediction of the syngas composition. They counseled the usage

of quasi-equilibrium temperature where the reaction of chemical is presumed to attain

equilibrium, rather than the real effective temperature of gasification unit. The value of

quasi-equilibrium temperature can be found through experimental statistics. This

technique is a conciliation of equilibrium thermodynamic models with experimental

models and does not need evidence on the dimensions, potential, and structure of the

gasifier. Doherty et al. (2009a) found the quasi-equilibrium technique for the minimization

of Gibb’s free energy with the help of the confined equilibrium technique for the

standardization of the model information against experimental facts, identifying a

temperature method for the reaction involved in gasification. Jarungthammachote and

Dutta (2008) put forth a modified equilibrium model based on the minimization of Gibbs

free energy at the equilibrium state. The model estimated the composition of six major

gases, CO, CO2, CH4, H2O, H2, and N2, in the producer gas. The composition of the

producer gas was determined through elemental balance rather than being constrained by
chemical reactions. The accuracy of the model was improved by including the effects of

carbon conversion; however, the model overestimated the concentration of hydrogen gas

for different configurations of spouted beds.

Subsequently, the original system scheme pronounced previously through the authors

(Moneti et al., 2016; Pallozzi et al., 2016) is changed by utilizing the change in the

pressure swing adsorption to a separation module together with palladium membranes so

that it will examine the feasible variant inside the normal performance, contemplating the

optimized pressure swing adsorption parameters that are presently applied in industry. The

water-gas shift method involves the separation of the products to acquire highly pure

hydrogen. The idea and application of palladium membranes for the departure of hydrogen Commented [CE2]: AU: Please check the sentence ‘The idea and
application of palladium membranes…’ for clarity and amend this if
required.
have lengthy been blanketed through the early work of Gryaznov, (1999) as well as

Rahimpour et al. (2017), representing the incredible selectivity of those membranes closer

to hydrogen, where the purity of hydrogen can be obtained up to 99.99% within the

permeate stream (Mahecha-Botero et al., 2008). In another study, Marcantonio et al.

(2019) simulated hazelnut shell gasification in a commercial process modeling software

Aspen plus for a fluidized bed gasifier with water-gas shift (WGS) and pressure swing

adsorption (PSA), through a quasi-equilibrium approach based on minimization of Gibb’s

free energy. The results were validated with the classic UNIFHY (UNIQUE gasifier for

Hydrogen production) model. The model was enhanced by replacing the PSA unit with a

palladium membrane, which improved the hydrogen recovery ratio from 38% to 49%. The

results were in good agreement with experimental data for the temperature of 785–870°C

and a steam-to-biomass (SB) ratio of 0.4–0.5.

In a dual fluidized bed, gasification of biomass and combustion of char appear in

two distinct fluidized beds, the heat required for the gasification of biomass is supplied via

combustion of char with the help of particles that recirculate. A dual fluidized consists of

various types of fluidized beds. Like, a device with combustion in a fast fluidized bed
(FFB) and gasification in a bubbling fluidized bed (BFB) is frequently applied (Liu et al.,

2015). But the layout and scale-up are difficult because of the complicated hydrodynamics

inside the system (Behzadi et al., 2019). Li et al. (2019) simulated hydrogen-rich syngas

production from biomass through different configurations of staged gasification, pyrolysis

gas reforming, and char combustion and studied the exergy analysis for different modes of

operation. The analysis showed that the generation of hydrogen from the hydrogen in

biomass is energetically more favorable. Pyrolysis gas combustion with char gasification

and staged gasification with pyrolysis gas reforming with char combustion have high

efficiency in the temperature range of 700–750°C with SB of 0.6–0.8, while staged

gasification with pyrolysis gas reforming and char gasification showed high efficiency in

the range of 650–700°C with SB of 0.7–0.8. CFD has been implemented to look at the

hydrodynamics in a dual fluidized bed. The Eulerian–Eulerian method is normally used

because it needs much less computational resources compared to the Eulerian–Lagrangian

technique. Nguyen et al. (2012) established a 2D Eulerian–Eulerian method with the aid of

evaluating with cold state tests gasification in a pilot-scale dual fluidized bed. The model

changed to look at the stable flow in the recycle loop and the consequence of the loop-seal

controller of the dual fluidized bed machine. The 3D model for the simulation of biomass

steam gasification in a dual fluidized bed reactor was developed by Yan et al. (2018).

Most of the present complete-loop CFD fashions use a single-drag model in

modeling a system inclusive of both a fast fluidized bed and a bubbling fluidized bed. As

an instance, the Gidaspow drag model was used to model both fast fluidized bed and

bubbling fluidized bed. But the model generally requires a minor length of the grid (e.g.,

20 times of particle length), which is computationally dense for the big-scale device, and it

usually overemphasizes the momentum change among gas and solid phase for immediate

fluidization (Lu et al., 2011). Heterogenous EMMS drag models are less grid-dependent

than that of the Gidaspow drag model and various so EMMS drag models are highly
suitable for the simulation of large-scale structures (Luo et al., 2017). Whereas because of

the distinctive structure of flow for fast fluidized bed and bubbling fluidized bed, distinct

EMMS drag models are required so that it will reap correct simulation consequences in a

dual bed system.

<H2>5.4.1 Computational Modeling Case Studies

The computational model can be executed with the help of commercial software such as

FLUENT, Barracuda, Star CCM+, and COMSOL. The detailed flowchart of the

computational work is described in Figure 5.1. The multi-phase simulation for the gas-

solid flow is performed by using different flow models. The mass and energy conversion

equations are satisfied for both the phases. The heat transfer takes place due to convection,

momentum transfer because of the drag force of gas and solid, and mass transfer takes

place due to heterogeneous reaction of chemicals.


Start

 Draw the geometry [2D/3D]


 Meshing with grid independency test

 Governing equation: Multi phase, energy, Species Transport


Equations
 Select the drag and viscous model

 Define the material properties for both phases


 Select the drag model, turbulent model

 Apply the appropriate boundary and initial conditions


 Select the appropriate solution method and solution control
 Initialize the solution method and run calculation for each
time step.

Solution is No
converged?

Yes

Observe the pattern of graph and


analyse the post processing

FIGURE 5.1 Flow path of modeling in different software.

The stages of the gasification process are drying, pyrolysis, partial combustion,

cracking, and reduction. These stages generally occur with an increase in temperature

inside the gasifier. The feed material is dried and exposed to a temperature around 300–
400°C. Then, the volatilization process of inventory starts, causing the molecule to dry.

Generally, this kind of reaction works at temperatures over 700°C (George et al., 2018;

Safarian et al., 2020). Subsequently, the drying and volatilization cycles might be

considered as happening immediately at the feed zone. When the drying and volatilization

measures have been finished, the particle of char starts burning and gasifying, and the

subsequent stages a homogeneous flow model. The following assumptions were Commented [CE3]: AU: Please check the text ‘….and the
subsequent stages a homogeneous flow model’ for clarity.

considered by the CFD solvers for the dynamics of fluid and chemical reactions while

solving the governing equations.

 The momentum transfer between the particles is not a function of the

temperature of the endothermic or exothermic reaction causing no effect on the

granular temperature.

 The segment of solid is thick and ceaseless in the riser and contact with the

wall of the gasifier; the mean free path of radiation is a lot more modest than

the solid molecule measurements to restrict the commitment of radiative

transfer of heat among the bed and wall of the gasifier. The gaseous molecules

can be expected to be transparent, so the energy due to radiation cannot be

absorbed or emitted. Therefore, the temperature of the gasifier will turn out to

be quickly uniform due to the solid fomentation of the particles. With these

expectations, it is sensible to consider the negligible loss of heat due to

radiative as compared to convective heat transfer. So, this is the main

component of heat transfer inside the gasifier.

 Feed materials are considered mono-scattered, smooth, inelastic, and

isothermal (negligible internal thermal resistance) spheres.

 Volatilization and drying are expected to occur instantaneously in the feeding

zone (Puig-Arnavat et al., 2013; Capizzi et al., 2020).

 Ash present in the feed material is assumed to be negligible.


<H2>5.4.2 Minimization of Gibbs Free Energy

There are two methods for the creation of Equilibrium Models. The first approach

is focused on the constants of equilibrium. This process needs to characterize the chemical

reaction used in the calculation (Loha et al., 2011). This signifies that proper substance

responses and data sets for proper harmony of constants are required. Reynolds Doherty et

al. (2009b) revealed the drawbacks associated with this approach for the creation of

equilibrium models. Researchers have suggested that this approach is not suitable for

complex modeling problems. Hence, the second model, i.e., minimization of the Gibbs

free energy is to be emphasized. The second model has the advantage that chemical

reaction kinetics is not required for the model solution.

Under the equilibrium condition, the overall Gibbs free energy of the framework is

minimized. The overall Gibbs free energy of a framework is characterized as

n
Gt   N i i (5.1)
i 1

where Gt is the overall Gibbs free energy and Ni is the number of moles of i species. µi is

the compound capable of species i that can be introduced by

 fi 
i  Gi  RT ln   (5.2)
 fo 

where R is gas constant and T is temperature, individually. fi addresses the fugacity of

species i and o signifies a standard thermodynamic amount, so Gi is standard Gibbs free

energy.

<H2>5.4.3 Energy Balance


With the increase in temperature different reactions are occurring in the gasifier

which can be categorized as endothermic and exothermic. Examples of such reactions are

the Boudouard reaction and partial oxidation. The heat produced in the exothermic

chemical reaction is devoured by the endothermic chemical reaction, and the remaining

heat is used as sensible heat that increases the temperature of the riser. The temperature of

the reaction is one of the most important parameters that influence the chemical reaction

as well as thermodynamic estimations. So, it is very important to know the temperature at

which the reaction takes place. The details of such reactions are presented in Table 5.3.

TABLE 5.3 The Main Reaction that Occurs Inside the Gasifier.

Reaction enthalpy
Reaction Eq. no
(kJ/mol)

Boudouard: C  CO2  2CO +172 (5.3)

Water-gas: C  H 2O  CO  H 2 +131 (5.4)

Hydrogasification: C  2 H 2  CH 4 –74.8 (5.5)

Partial oxidation: C  0.5 O2  CO –111 (5.6)

C  O2  CO2 –394 (5.7)

CO  0.5O2  CO2 –284 (5.8)


Oxidation:
CH 4  2O2  CO2  2 H 2O –803 (5.9)

H 2  0.5O2  H 2 O –242 (5.10)

Shift conversion: CO  H 2O  CO2  H 2 –42 (5.11)

2CO  2 H 2  CH 4  CO2 –247 (5.12)

Methanation: CO  3 H 2  CH 4  H 2O –206 (5.13)

CO2  4 H 2  CH 4  2 H 2O –165 (5.14)


CH 4  H 2O  CO  3H 2 +206 (5.15)
Steam reforming:
CH 4  0.5O2  CO  2 H 2 –36 (5.16)

m (5.17)
Tar cracking: Cn Hm +n CO2  +2n CO –
2H2

Char reaction: Char +O2  CO2 +CO – (5.18)

To determine the temperature, 1st law of thermodynamics or energy balance is

implemented for the gasification

Qloss   nr H rTr   n p H pT p  H (5.19)


r p

where Qloss is the loss of heat in the gasification cycle. Heat loss is almost 1% of the HHV

of the inventory (Pallozzi et al., 2016), so in Eq. (5.19) in place of heat loss, 1% of HHV

value can be used. H is the value of enthalpy while suffix r and p denote the reactant and

product respectively at their specific temperature. The value of enthalpy can be calculated

by

T
H  (T )  H f (298)  C
298
p (T) dT (5.20)

For these estimations, the reactants are assumed to be at room temperature that is

298 K, and hence, the formation of enthalpy with N2 and O2 is zero. Enthalpy can also be

calculated with the relation proposed by Jarungthammachote and Dutta, (2008).

H f , fuel LHV   nk ( H f ) k  (5.21)


k

Here, Hf is the formation of enthalpy for the kth component at complete

combustion of feedstock, and LHV is the lower heating value of the feedstock. Higher

heating value (HHV) can be calculated as


HHV  0.349C  1.178H  0.1S  0.1O  0.15N  0.02Ash (5.22)

The lower heating value of the feed material can also be calculated by

LHV  HHV  9mH (h fg ) (5.23)

where C, H, O, N, and S are the rates of mass of carbon, hydrogen, oxygen, nitrogen, and

sulfur in the feedstock. The mH is the fractional mass of hydrogen in feedstock, hfg is the

enthalpy of vaporization of water. For estimating the temperature of the chemical reaction,

the initial temperature is predicted, and it is utilized to estimate the value of the producer

gas in place of Gibbs free energy minimization. The expected temperature and produced

gas segments are subbed to process the energy balance.

<H1>5.5 CASE STUDIES FOR HYDROGEN PRODUCTION

Figures 5.2 and 5.3 show the effect of pressure on the composition of syngas for

two different inventories, which is the content of H2, and CH4. The concentration of H2

decreases from 7.62 to 6.75% for the sugarcane bagasse, while for sawdust it decreases

from 9.24 to 8.51% with the enhancement of pressure from 1 to 4 bar. The concentration

of H2 decreases with an enhancement of pressure mainly due to a higher pressure

Methanation reaction enhanced. Methanation reaction converts the molecules of hydrogen Commented [CE4]: AU: Please check the text ‘….a higher
pressure Methanation reaction enhanced’ for clarity.

into methane after reacting with carbon that is present in the feed material. It can also be

observed that the composition of CH4 increases with pressure enhancement. At higher

pressure, Methanation reaction enhances the production of CH4 from 3.16 to 3.96% for

sugarcane bagasse, while the sawdust percentage of methane increases 3.8 to 5.2% which

is around 37% enhancement due to increase of pressure from 1 to 4 bar.


Sawdust
10
Gas composition %

8
6
4 H2
2 CH4
0
1 2 3 4
Pressure (bar)

FIGURE 5.2 Effect of pressure on gas composition for sawdust.

Sugarcane bagasse
10
Gas composition %

8
6
4
H2
2
0
1 2 3 4
Pressure (bar)

FIGURE 5.3 Effect of pressure on gas composition for sugarcane bagasse.

<H1>5.6 CONCLUSIONS

As a clean fuel as well as for high energy hydrogen has many benefits. Among the

different sources accessible for the production of hydrogen, gasification of biomass has

many advantages over another capability mainly due to availability of biomass in

abundance, furthermore giving CO2-impartial fuel. When gasification of biomass is done

in the presence of a steam medium for the production of hydrogen, its partial pressure of

hydrogen is highest in the composition of syngas.

The work that has been done for the simulation and modeling of biomass

gasification showed that various works have been published based on kinetic
demonstrating of gasifier; however, exceptionally restricted literatures are available on

biomass steam gasification for the production of hydrogen. Besides, only equilibrium

models were established for the production of hydrogen dependent on steam gasification

with in situ carbon dioxide capture. There is a need to grow more comprehensive models

which incorporate the carbonation response at the same time.

Various forms of flowsheets are accessible for the production of hydrogen through

biomass. It is important to create an improved flowsheet with a minimum number of units

and pretend enhanced production of hydrogen from various biomass. Moreover, such a

flowsheet model ought to have the option to examine the performance of the gasifier at the

same time as the kinetic model.

The greater part of the monetary examinations for hydrogen creation from biomass

gasification was completed utilizing the decent qualities of the capital venture, utilities,

and at explicit working boundaries. It is prescribed for future exploration work to foster a

model which can figure the expense of hydrogen creation by settling the flowsheet

boundaries and different factors which influence the production of hydrogen cost.

Efficient plan calculations for the production of hydrogen from biomass steam gasification

are infrequently announced. There is a need to introduce a comprehensive framework that

ought to incorporate imitation of an economic flowsheet, the heat required for the

interaction, and complete financial investigation for the production of hydrogen from

different available biomass feedstocks.

<H1>REFERENCES

Abdoulmoumine, N.; Kulkarni, A.; Adhikari, S.Effects of Temperature and Equivalence

Ratio on Pine Syngas Primary Gases and Contaminants in a Bench-scale Fluidized Bed

Gasifier. Ind. Eng. Chem. Res. 2014, 53 (14), 5767–5777.


Adeyemi, I.; Janajreh, I.; Arink, T.; Ghenai, C.Gasification Behavior of Coal and Woody

Biomass: Validation and Parametrical Study. Appl. Energy 2017, 185, 1007–1018.

Ardila, Y. C.; Figueroa, J. E. J.; Lunelli, B. H.; Filho, R. M.; Maciel, M. R. W.Syngas

Production From Sugar Cane Bagasse in a Circulating Fluidized Bed Gasifier Using

Aspen PlusTM: Modelling and Simulation. Comput. Aided. Chem. Eng. 2012, 30 (June),

1093–1097.

Barba, D.; Capocelli, M.; Cornacchia, G.; Matera, D. A.Theoretical and Experimental

Procedure for Scaling-up RDF gasifiers: The Gibbs Gradient Method. Fuel [Internet].

2016, 179, 60–70.

Basu, P.; Mettanant, V.Biomass Gasification in Supercritical Water -- A Review. Int. J.

Chem. React. Eng. 2009, 7 (1).

Behzadi, A.; Habibollahzade, A.; Ahmadi, P.; Gholamian, E.; Houshfar, E.Multi-

Objective Design Optimization of a Solar Based System for Electricity, Cooling, and

Hydrogen Production. Energy 2019, 169, 696–709.

Berrueco, C.; Montané, D.; Matas Güell, B.; del Alamo, G.Effect of Temperature and

Dolomite on tar Formation During Gasification of Torrefied Biomass in a Pressurized

Fluidized Bed. Energy 2014, 66, 849–859.

Berrueco, C.; Recari, J.; Güell, B. M.; del Alamo G. Pressurized Gasification of Torrefied

Woody Biomass in a Lab Scale Fluidized Bed. Energy 2014, 70, 68–78.

Capareda, S.; Powell, J.; Aquino, F.Performance of a Portable Downdraft Gasifier. 2007

Beltwide Cott. Conf. 2007, 1500–1504.

Capizzi, G.; Sciuto, G. L.; Napoli, C.; Woźniak, M.; Susi, G.A Spiking Neural Network-

based Long-term Prediction System for Biogas Production. Neural Netw. 2020, 129, 271–

279.
Dogru, M.; Howarth, C. R.; Akay, G.; Keskinler, B.; Malik, A. A.Gasification of Hazelnut

Shells in a Downdraft Gasifier. Energy 2002, 27 (5), 415–427.

Doherty, W.; Reynolds, A.; Kennedy, D.The Effect of Air Preheating in a Biomass CFB

Gasifier Using ASPEN Plus Simulation. Biomass Bioenergy 2009a, 33 (9), 1158–1167.

Doherty, W.; Reynolds, A.; Kennedy, D.The effect of Air Preheating in a Biomass CFB

Gasifier Using ASPEN Plus Simulation. Biomass Bioenergy 2009b, 33 (9), 1158–1167.

Duan, F.; Jin, B.; Huang, Y.; Li, B.; Wu, Y.; Zhang, M.Results of Bituminous Coal

Gasification Upon Exposure to a Pressurized Pilot-plant Circulating Fluidized-bed (CFB)

reactor. Energ. Fuel. 2010, 24 (5), 3150–3158.

Fermoso, J.; Arias, B.; Plaza, M. G.; Pevida, C.; Rubiera, F.; Pis, J. J.; García-Peña, F.;

Casero, P.High-pressure Co-gasification of Coal with Biomass and Petroleum Coke. Fuel

Process. Technol. 2009, 90 (7–8), 926–932.

Figueroa, J. E. J.; Ardila, Y. C.; Filho, R. M.; Maciel, M. R. W.Fluidized Bed Reactor for

Gasification of Sugarcane Bagasse: Distribution of Syngas, Bio-Tar and Char. Chem. Eng.

Trans. 2014, 37, 229–234.

Frigo, S.; Spazzafumo, G.Cogeneration of power and substitute of natural gas using

biomass and electrolytic hydrogen. Int. J. Hydrogen. Energy. 2018, 43 (26), 11696–11705.

García-Ibañez, P.; Cabanillas, A.; Sánchez, J. M.Gasification of Leached Orujillo (Olive

Oil Waste) in a Pilot Plant Circulating Fluidised Bed Reactor. Preliminary results.

Biomass Bioenergy. 2004, 27 (2), 183–194.

George, J.; Arun, P.; Muraleedharan, C.Assessment of Producer Gas Composition in air

Gasification of Biomass Using Artificial Neural Network Model. Int. J. Hydrogen. Energy

2018, 43 (20), 9558–9568.

Gryaznov, V.Membrane Catalysis. Catal. Today 1999, 51 (3–4), 391–395.


Guan, J.; Wang, Q.; Li, X.; Luo, Z.; Cen, K.Thermodynamic Analysis of a Biomass

Anaerobic Gasification Process for Hydrogen Production with Sufficient CaO. Renew.

Energy 2007, 32 (15), 2502–2515.

Gül, S.; Akgün, F.; Aydar, E.; Ünlü, N.Pressurized Gasification of Lignite in a Pilot Scale

Bubbling Fluidized Bed Reactor With Air, Oxygen, Steam and CO2 Agents. Appl. Therm.

Eng. 2018, 130, 203–210.

Han, L.; Wang, Q.; Luo, Z.; Rong, N.; Deng, G.H2 rich Gas Production via Pressurized

Fluidized Bed Gasification of Sawdust with In Situ CO2 Capture. Appl. Energy 2013, 109,

36–43.

Huang, J.; Fang, Y.; Chen, H.; Wang, Y.Coal Gasification Characteristic in a Pressurized

Fluidized Bed. Energ. Fuel. 2003, 17 (6), 1474–1479.

Jarungthammachote, S.; Dutta, A.Equilibrium Modeling of Gasification: Gibbs Free

Energy Minimization Approach and Its Application to Spouted Bed And Spout-Fluid Bed

Gasifiers. Energy Convers Manag. 2008, 49 (6), 1345–1356.

Jong, W. D.; Andries, J.; Hem, K. R. G.No Title. 1999, 16, 0–3.

Jong, W. D.; Hein, K. R. G.Nitrogen Compounds in Pressurised Fluidised Bed

Gasification of Biomass and Fossil Fuels; 2005.

Kamińska-Pietrzak, N.; Smoliński, A.Selected Environmental Aspects of Gasification and

Co-Gasification of Various Types of Waste. J. Sustain. Min. 2013, 12 (4), 6–13.

Karmakar, M. K.; Datta, A. B.Generation of Hydrogen Rich Gas Through Fluidized Bed

Gasification of Biomass. Bioresour. Technol. 2011, 102 (2), 1907–1913.

Key World Energy Statistics 2018 [Online],2018. https://www.oecd-

ilibrary.org/energy/key-world-energy-statistics-2018_key_energ_stat-2018-en

Kitzler, H.; Pfeifer, C.; Hofbauer, H.Pressurized Gasification of Woody Biomass-


Variation of Parameter. Fuel. Process. Technol. 2011, 92 (5), 908–914.

Klimanek, A.; Bigda, J.CFD Modelling of CO2 Enhanced Gasification of Coal in a

Pressurized Circulating Fluidized Bed Reactor. Energy 2018, 160, 710–719.

Konstantinou, E.; Marsh, R.Experimental Study on the Impact of Reactant Gas Pressure in

the Conversion of Coal Char to Combustible Gas Products in the Context of Underground

Coal Gasification. Fuel 2015, 159, 508–518.

Kurkela, E.; Kurkela, M.; Moilanen, A.; Fin, W. Fluidised-bed Gasification of High-alkali

Biomass Fuels.

Li, G.; Liu, Z.; Liu, F.; Yang, B.; Ma, S.; Weng, Y.; Zhang, Y.; Fang, Y.Advanced Exergy

Analysis of Ash Agglomerating Fluidized Bed Gasification. Energy. Convers. Manag.

2019, 199 (2001).

Li, K.; Zhang, R.; Bi, J.Experimental Study on Syngas Production by Co-gasification of

Coal and Biomass in a Fluidized Bed. Int. J. Hydrogen Energy 2010, 35 (7), 2722–2726.

Li, Q.; Song, G.; Xiao, J.; Sun, T.; Yang, K.Exergy Analysis of Biomass Staged-

Gasification for Hydrogen-Rich Syngas. Int. J. Hydrogen Energy 2019, 44 (5), 2569–

2579.

Li, Q.; Zhang, M.; Zhong, W.; Wang, X.; Xiao, R.; Jin, B.Simulation of Coal Gasification

in a Pressurized Spout-Fluid Bed Gasifier. Can. J. Chem. Eng. 2009, 87 (2), 169–176.

Li, X.; Grace, J. R.; Watkinson, A. P.; Lim, C. J.; Ergüdenler, A.Equilibrium Modeling of

Gasification: A Free Energy Minimization Approach and Its Application to a Circulating

Fluidized bed Coal Gasifier. Fuel 2001, 80 (2), 195–207.

Liu, H.; Cattolica, R. J.; Seiser, R.; Liao C hsien.Three-dimensional Full-loop Simulation

of a Dual Fluidized-bed Biomass Gasifier. Appl. Energy 2015, 160, 489–501.

Loha, C.; Chatterjee, P. K.; Chattopadhyay, H.Performance of Fluidized Bed Steam


Gasification of Biomass—Modeling and Experiment. Energy Conv. Manag. 2011, 52 (3),

1583–1588.

Lu, B.; Wang, W.; Li, J.Eulerian Simulation of Gas-solid Flows with Particles of Geldart

Groups, A.; B and D using EMMS-based Meso-scale Model. Chem. Eng. Sci. 2011, 66

(20), 4624–4635.

Luo, H.; Lu, B.; Zhang, J.; Wu, H.; Wang, W.A grid-independent EMMS/bubbling Drag

Model for Bubbling and Turbulent Fluidization. Chem. Eng. J. 2017, 326, 47–57.

Mahecha-Botero, A.; Boyd, T.; Gulamhusein, A.; Comyn, N.; Lim, C. J.; Grace, J. R.;

Shirasaki, Y.; Yasuda, I.Pure Hydrogen Generation in a Fluidized-bed Membrane Reactor:

Experimental Findings. Chem. Eng. Sci. 2008, 63 (10), 2752–2762.

Mahmoudi, S.; Baeyens, J.; Seville, J.The solids flow in the CFB-riser Quantified by

Single Radioactive Particle Tracking. Powder Technol. 2011, 211 (1), 135–143.

Mallick, D.; Mahanta, P.; Moholkar, V. S.Co–gasification of Coal/biomass Blends in 50

kWe Circulating Fluidized Bed Gasifier. J. Energy Inst. 2019, (xxxx).

Mallick, D.; Mahanta, P.; Moholkar, V. S.Co–gasification of coal/biomass Blends in 50

kWe Circulating Fluidized Bed Gasifier. J. Energy Inst. 2020, 93 (1), 99–111.

Marcantonio, V.; De Falco, M.; Capocelli, M.; Bocci, E.; Colantoni, A.; Villarini,

M.Process Analysis of Hydrogen Production from Biomass Gasification in Fluidized Bed

Reactor with Different Separation Systems. Int. J. Hydrogen Energy. 2019, 44 (21),

10350–10360.

Mayerhofer, M.; Mitsakis, P.; Meng, X.; De Jong, W.; Spliethoff, H.; Gaderer,

M.Influence of Pressure, Temperature and Steam on tar and Gas in Allothermal Fluidized

Bed Gasification. Fuel 2012, 99, 204–209.

Moneti, M.; Di Carlo, A.; Bocci, E.; Foscolo, P. U.; Villarini, M.; Carlini, M.Influence of
the Main Gasifier Parameters on a Real System for Hydrogen Production From Biomass.

Int. J. Hydrogen Energy 2016, 41 (28), 11965–11973.

Motta, I. L.; Miranda, N. T.; Maciel Filho, R.; Wolf Maciel, M. R.Sugarcane Bagasse

Gasification: Simulation and Analysis of Different Operating Parameters, fluidizing

media, and gasifier types. Biomass Bioenergy. 2019, 122 (March 2018), 433–445.

Nguyen, T. D. B.; Seo, M. W.; Lim, Y. Il.; Song, B. H.; Kim, S. D.CFD Simulation with

Experiments in a Dual Circulating Fluidized Bed Gasifier. Comput. Chem. Eng. 2012, 36

(1), 48–56.

Padban, N.; Wang, W.; Ye, Z.; Bjerle, I.; Odenbrand, I.Tar Formation in Pressurized

Fluidized Bed Air Gasification of Woody Biomass. Energ. Fuel. 2000, 14 (3), 603–611.

Pallozzi, V.; Di Carlo, A.; Bocci, E.; Villarini, M.; Foscolo, P. U.; Carlini, M.Performance

Evaluation at Different Process Parameters of an Innovative Prototype of Biomass

Gasification System Aimed to Hydrogen Production. Energy Conv. Manag. 2016, 130,

34–43.

Pellegrini, L. F.;, de Oliveira, S.Exergy Analysis of Sugarcane Bagasse Gasification.

Energy 2007, 32 (4), 314–327.

Pinto, F.; André, R. N.; Franco, C.; Lopes, H.; Gulyurtlu, I.; Cabrita, I.Co-gasification of

Coal and Wastes in a Pilot-scale Installation 1: Effect of Catalysts in Syngas Treatment to

Achieve Tar Abatement. Fuel 2009, 88 (12), 2392–2402.

Puig-Arnavat, M.; Hernández, J. A.; Bruno, J. C.; Coronas, A.Artificial Neural Network

Models for Biomass Gasification In Fluidized Bed Gasifiers. Biomass Bioenergy. 2013,

49:279–289.

Qin, Y. H.; Huang, H. F.; Wu, Z. B.; Feng, J.; Li, W.; Xie, K. C.Characterization of Tar

from Sawdust Gasified in the Pressurized Fluidized Bed. Biomass Bioenergy 2007, 31 (4),
243–249.

Rahimpour, M. R.; Samimi, F.; Babapoor, A.; Tohidian, T.; Mohebi, S.Palladium

Membranes Applications in Reaction Systems for Hydrogen Separation and Purification:

A Review. Chem. Eng. Process. Process. Intensif. 2017, 121 (February), 24–49.

Safarian, S.; Ebrahimi Saryazdi, S. M.; Unnthorsson, R.; Richter, C.Artificial Neural

Network Integrated with Thermodynamic Equilibrium Modeling of Downdraft Biomass

Gasification-power Production Plant. Energy 2020, 213, 118800.

Safarian, S.; Unnthorsson, R.; Richter, C.Simulation and Performance Analysis of

Integrated Gasification-Syngas Fermentation Plant for Lignocellulosic Ethanol

Production. Fermentation 2020, 6 (3).

Sahoo, A.; Ram, D. K.Gasifier Performance and Energy Analysis for Fluidized Bed

Gasification of Sugarcane Bagasse. Energy 2015, 90,1420–1425.

Sánchez, C.; Arenas, E.; Chejne, F.; Londoño, C. A.; Cisneros, S.; Quintana, J. C.A New

Model for Coal Gasification on Pressurized Bubbling Fluidized Bed Gasifiers. Energy

Conv. Manag. 2016, 126, 717–723.

Seo, M. W.; Goo, J. H.; Kim, S. D.; Lee, S. H.; Choi, Y. C.Gasification Characteristics of

Coal/biomass Blend in a Dual Circulating Fluidized Bed Reactor. Energ. Fuel. 2010, 24

(5), 3108–3118.

Sharma, A. K.Experimental Investigations on a 20 kWe, Solid Biomass Gasification

System. Biomass Bioenergy. 2011, 35 (1), 421–428.

Shayan, E.; Zare, V.; Mirzaee, I.Hydrogen Production from Biomass Gasification; A

Theoretical Comparison of Using Different Gasification Agents. Energy Conv. Manag.

2018, 159 (August 2017), 30–41.

Sheth, P. N.; Babu B, V.Experimental Studies on Producer Gas Generation from Wood
Waste in a Downdraft Biomass Gasifier. Bioresour. Technol. 2009, 100 (12), 3127–3133.

Srinivas, T.; Gupta, S.; Reddy, B. V.Thermodynamic Equilibrium Model and Exergy

Analysis of a Biomass Gasifier. J. Energy Resour. Technol. 2009, 131 (3), 031801.

Srinivas, T.; Gupta, S.; Reddy, B. V.; Nag, P. K.Parametric Analysis of a Coal Based

Combined Cycle Power Plant. Int. J. Energy Res. 2006, 30 (1), 19–36.

Suksuwan, W.; Wae, M.; Mel, M.Akademia Baru Journal of Advanced Research in Fluid

Development of Mini Pilot Fluidized Bed Gasifier for Industrial Approach : Preliminary

Study Based on Continuous Operation Akademia Baru Combust Fine Solid Fuels Having

Non-uniform Size, is Applied in 2018, 1 (1), 35–43.

Valin, S.; Ravel, S.; Guillaudeau, J.; Thiery, S.Comprehensive Study of the Influence of

Total Pressure on Products Yields in Fluidized Bed Gasification of Wood Sawdust. Fuel.

Process. Technol. 2010, 91 (10), 1222–1228.

Wu, C.; Williams, P. T.Pyrolysis-gasification of Plastics, Mixed Plastics and Real-World

Plastic Waste with and Without Ni-Mg-Al Catalyst. Fuel 2010, 89 (10), 3022–3032.

Xiao, R.; Zhang, M.; Jin, B.; Huang, Y.; Zhou, H.High-Temperature Air/steam-Blown

Gasification of Coal in a Pressurized Spout-Fluid Bed. Energ. Fuel. 2006, 20 (2), 715–

720.

Xiao, R.; Zhang, M.; Jin, B.; Huang, Y.; Zhou, H.; Motta, I. L.; Miranda, N. T.; Maciel

Filho, R.; Wolf Maciel, M. R.; Li, J.; et al.Equilibrium Modeling of Gasification: A Free

Energy Minimization Approach and Its Application to a Circulating Fluidized Bed Coal

Gasifier. Biomass Bioenergy 2018, 53 (2), 433–445.

http://dx.doi.org/10.1016/j.fuel.2012.04.022

Yan, L.; Cao, Y.; Zhou, H.; He, B.Investigation on Biomass Steam Gasification in a Dual

Fluidized Bed Reactor with the Granular Kinetic Theory. Bioresour. Technol. 2018, 269
(July), 384–392.

Zeng, J.; Xiao, R.; Zeng, D.; Zhao, Y.; Zhang, H.; Shen, D.High H2/CO Ratio Syngas

Production from Chemical Looping Gasification of Sawdust in a Dual Fluidized Bed

Gasifier. Energ. Fuel. 2016, 30 (3), 1764–1770.

Żogała, A.Equilibrium Simulations of Coal Gasification – Factors Affecting Syngas

Composition. J. Sustain. Min. 2014, 13 (2), 30–38.

You might also like