You are on page 1of 15

Powder Technology 296 (2016) 87–101

Contents lists available at ScienceDirect

Powder Technology
journal homepage: www.elsevier.com/locate/powtec

Biomass gasification with CO2 in a fluidized bed


Yongpan Cheng a,b, Zhihao Thow a, Chi-Hwa Wang a,b,⁎
a
Department of Chemical and Biomolecular Engineering, National University of Singapore, 4 Engineering Drive 4, 117585, Singapore
b
NUS Environmental Research Institute, National University of Singapore, 5A Engineering Drive 1, #02-01, 117411, Singapore

a r t i c l e i n f o a b s t r a c t

Article history: Biomass gasification with CO2 as the gasifying agent is a promising way to relieve the energy shortage and min-
Received 3 November 2014 imize CO2 emission. In this study, Eulerian method was used to study the CO2 gasification of biomass in a fluidized
Received in revised form 18 December 2014 bed gasifier. It was found that the CO2 percentage in the gasifying mixture with air, the CO2-to-biomass ratio, the
Accepted 22 December 2014
moisture content of woodchips, and the woodchip size had great influence on the gasification performance. With
Available online 27 December 2014
the increasing CO2-to-biomass ratio, the mole fraction of CO in the producer gases would be increased while
Keywords:
those of H2 and CO2 vary in the opposite trend. When CO2 mass percentage in the gasifying agent was 60%,
CO2 capture the fractions of CO and CH4 in the producer gas reached the maximum, as well as the lower heating value and
Biomass gasification cold gas efficiency, so this was the optimal condition when the gasifier had the best performance. The moisture
Renewable energy content and particle size of the woodchips had negative effects on the gasification performance, because of the
Numerical simulation lower heating value of producer gases, cold gas efficiency and CO2 conversion ratio were both reduced with
increasing moisture content and particle size. This study offers a promising way to integrate the gasification of
renewable biomass with CO2 capture, and may be helpful in the design and operation of biomass gasifier.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction In the literature, there are many studies on the topic of biomass
utilization, covering various subjects such as fermentation, combustion
The explosive increasing energy consumption is one of the critical and gasification. Fermentation does not have high requirement for the
challenges throughout the world, and currently significant percentage feedstock, such as limitation on the moisture content, but the process
of the consumed energy comes from fossil fuels, such as petroleum, is quite slow and usually need large reactors to ensure significant output
coal and natural gases. According to Song [1], the total energy consump- of producer gas. Comparatively, combustions are fast processes, so the
tion in the 20th century was about 10,048 million tons of oil equivalent, reactors can be quite compact. However, the exhaust gases from bio-
with 24% from coal, 39% from petroleum, 23% from natural gas, 6% from mass combustion can include dust, NOx, SOx or heavy metals etc,
nuclear power, and only 8% from renewable energy, including hydro- which are costly to be removed. Therefore, compared with fermentation
electric power, biomass, geothermal, solar and wind energy. On one and combustion, gasification is an effective and clean way to convert
hand, the limited and non-renewable fossil fuels have been consumed biomass into useful fuels and chemical feedstocks [3,4]. With proper
rapidly and will be depleted in the near future; on the other hand, ener- cleaning of the fuels produced, they can be directly used in electricity
gy generation from fossil fuels is also the major source for CO2 emission and heat production devices, such as internal combustion engines, gas
(about 41%) [2], which is the major greenhouse gas contributing to turbines and fuel cells [5,6]. So far there have been a great amount of
global warming. Furthermore, the emission of NOx and SOx can result reviews on the study of biomass [7–14], however, biomass gasification
in acid rain, which is also a great threat to the environment. In order with CO2 as gasifying agent is seldom addressed.
to resolve the energy shortage and relieve global warming, biomass is Gasification with CO2 has several advantages [15], for example,
considered a potential new and clean energy source. Biomass is a CO2 no energy is required for vaporization; the H2/CO ratio in producer
neutral and environmentally friendly energy source, as it is formed by gases can be easily adjusted to meet the specific requirement; CO2 can
the plant photosynthesis process, which absorbs CO2 from the atmo- produce more volatiles in a reactive char, so the gasification efficiency
sphere. Furthermore, biomass can be converted into gaseous, liquid can be improved. Finally the gasification of CO2 instead of nitrogen
and solid fuels, so it is convenient for storage and transportation. can lead to flue gas with high percentage of CO2, which is suitable for
direct recovery and recycle of CO2. CO2 has been successfully used to
gasify coal to produce syngas and relieve CO2 pollution [16–18]. The
thermochemical processes involved in biomass gasification with CO2
⁎ Corresponding author at: Department of Chemical and Biomolecular Engineering,
National University of Singapore, 4 Engineering Drive 4, 117585, Singapore. Tel.: +65
are similar to those in coal gasification with CO2. As the reaction of
65165079; fax: +65 67791936. CO2 with carbon is highly endothermic, and highly energy-intensive,
E-mail address: chewch@nus.edu.sg (C.-H. Wang). usually the mixture of CO2 and O2 or CO2 and steam is used as the

http://dx.doi.org/10.1016/j.powtec.2014.12.041
0032-5910/© 2014 Elsevier B.V. All rights reserved.
88 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

gasifying agents. Renganathat et al. [19] carried out a thermodynamic operated at the Institute of Energy Engineering at Berlin Institute of
analysis on CO2 utilization for gasification of carbonaceous feedstocks Technology [24], as shown in Fig. 1. The freeboard zone and the bub-
using Gibbs minimization approach, and found that when CO2 was com- bling fluidized bed zone have inner diameters of 0.135 m and 0.095 m
bined with steam or oxygen as a gasifying agent, the requirement for respectively. The gasifying agent (mixture of air and CO2) was intro-
carbon dioxide and energy could be reduced, as well as the carbon diox- duced at the bottom of gasifier, while the woods were fed into the gas-
ide conversion. Furthermore, the ratio of hydrogen/carbon in syngas ifier through a fuel inlet with diameter of 0.05 m at 0.08 m above the
could be varied in a wide range. Irfan et al. [16] reviewed the effect of bottom. The producer gases escaped from the outlet with diameter
different parameters on coal-char gasification in CO2 stream, such as 0.03 m at 0.05 m below the top. Instead of inert bed materials, char
the effect of coal rank, pressure, temperature, gas composition, catalyst was selected as bed materials, as done by [24]. The use of char as bed
and the minerals inside the coal, heating rate, particle size and reactor materials had some potential advantages, for example, the char had
types. As the extension of the study, they also reviewed the kinetics the capability to decompose tar [25,26], and it did not need to be regen-
and reaction rate equations for coal-char gasification in low tempera- erated in that char was a byproduct of biomass gasification, the pressure
ture and high temperature regions under low and high pressures. loss in the reactor was also lower due to its low density compared with
Mani et al. [20] studied the reaction kinetics and mass transfer of traditional bed materials. The wood and char had the constant diameter
wheat straw char gasification with CO2 through thermogravimetric of 2 mm and 4 mm, and had the density 605 kg/m3 and 450 kg/m3,
apparatus (TGA); the effects of temperature and particle size on the respectively. The gasifier was initially filled to a char bed height of
diffusion and surface reactions were identified. Also by virtue of TGA, 350 mm with volume fraction of 0.63. In order to quickly activate
Butterman and Castaldi [15] examined the gas evolution, mass decay the biomass gasification, the initial temperature of the char bed and
behavior, energy contents of several woods, grasses and agricultural gas was at 1000 K. The preheated gasifying agent with different
residues using a mixture of CO2 and steam with different proportions. compositions of air and CO2 was continuously introduced at a 970 K
Garcia et al. [21] carried out the catalytic CO2 gasification of pine saw- from the bottom of the gasifier during the entire operation time,
dust at a relatively low temperature and atmospheric pressure; the in- the wall of the gasifier was set as 970 K as well. 10 kg/h of wood at
fluence of the catalyst weight/biomass flow rate ratio was analyzed on 623 K was fed into the gasifier through fuel inlet under atmospheric
product distribution and gas composition. pressure.
As fluidized bed gasifier has high reaction rates and effective mixing
inside, it is widely used in medium and large scale biomass gasification. 2.2. Governing equations
Oevemann et al. [22] simulated wood gasification in a two-dimensional
bubbling fluidized bed reactor with an Eulerian–Lagrangian approach. In this study, due to the high solid fraction in the bubbling fluidized
The influence of wood feeding rate on the gasification was studied. bed, the Eulerian method was used to simulate the biomass gasification.
With a similar method, Xie et al. [23] simulated the gasification of for- The solid char was considered as the continuum, and the solid fluctuat-
estry residues in a three-dimensional fluidized bed. Their numerical ing energy was described with granular temperature from the kinetic
model could predict the product gas composition and carbon conver- theory [27]. The phases were able to interpenetrate into each other,
sion efficiency, in good agreement with experimental data. In addition, and the sum of all the volume mass fraction was unity. The accumula-
the flow regimes, profiles of particle species, and distribution of gas tion of mass in each phase was balanced by the convective mass fluxes.
compositions inside the reactor were also discussed. Gerber et al. [24] The biomass gasification included the fluid flow, heat and mass transfer,
modeled wood gasification in a bubbling fluidized bed reactor using as well as the chemical reactions; they were governed by the following
char as bed material with an Eulerian method. The product gas concen- equations.
tration and temperature were investigated under different operating
conditions and model parameters. Blasi [12] reviewed modeling in 2.2.1. Gas phase
chemical and physical processes of wood and biomass pyrolysis, espe- The conservation equation for the mass fraction of the species i in the
cially the chemical kinetics in primary reactions described by one- and gas phase could be written as [24]:
multi-component mechanisms, and secondary reactions of tar cracking
and polymerization. Gomez-Barea and Leckner [11] reviewed the
modeling of biomass gasification in fluidized bed. The mathematical
i
∂εg ρg Y g   Xn
! i  
þ ∇  ε g ρg v g Y g ¼ mi;g þ s
m : ð1Þ
reactor models for biomass and waste gasification in fluidized bed ∂t s¼1 i;s

were presented.
As biomass gasification experiments are usually quite expensive and
time-consuming, numerical simulation provides an efficient alternative With additional momentum terms between gas phase and solid
way to carry out such studies for providing guidance to design the gas- phase, the gas phase momentum equation could be written as:
ifier and optimize the experiments. In this study, the Eulerian method
will be used to study CO2 gasification of biomass in a fluidized bed. !
∂εg ρg v g   Xn !
The effects of compositions of CO2 and air in gasifying mixture, moisture !! !
þ ∇ εg ρg v g v g ¼ −εg ∇p−εg ∇  τg − s¼1
s
I sg −εg ρg g : ð2Þ
content of biomass, and particle size on the biomass gasification will be ∂t
studied in details. The gasification performance will be evaluated in
terms of gas composition and temperature, axial profiles of gas species,
lower heating value of the producer gases, cold-gas efficiency and frac- The momentum exchange term between solid phases and gas
tional CO2 conversion. This study integrates biomass gasification with phases Isg could be expressed as:
CO2 captures, which cannot only increase the power supply of renew-
able energy, but also reduce CO2 emissions effectively. !    
! ! ! !
I sg ¼ K sg v s − v g þ msg v s − v g : ð3Þ
2. Mathematical formulation

2.1. Physical model The dominant force in the gas and solid phase momentum balances
was the inter-phase momentum transfer, which was represented as
The numerical simulation of biomass gasification in this study was drag force; here the Syamlal–O'Brien drag model was used for calculat-
based on the lab-scale bubbling fluidized bed gasifier, which was ing the fluid–solid drag force [28].
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 89

Fig. 1. Schematic of fluidized bed for biomass gasification and computational grid.

The energy equation in gas phase could be expressed as: The transport equation for incident radiation, G, could be written as
[30]:
∂εg ρg Hg  
!
þ ∇  εg ρg v g Hg ð4Þ
∂t  
! Xn h   i 
−1 4
∇  ð3ðα þ σ s Þ−Cσ s Þ ∇G −αG þ 4ασT ¼ SG : ð9Þ
¼ −∇  q g þ s
s¼1
hgs T s −T g −msg hsg −ΔHg :

The total enthalpy of the gas phase, Hg was defined as follows: 2.3. Reaction model
"Z #
X X Tg   2.3.1. Pyrolysis
i i 0
Hg ¼ i
Y g Hi ¼ i
Yg C p;i dTþHi T re f ;i : ð5Þ In this model, three kinds of chemical processes were considered:
T re f
(1) pyrolysis of wood, (2) homogeneous gas phase reactions and
(3) heterogeneous char reactions. The reaction model included a system
of three phases, which were one gas phase and two solid phases (wood
2.2.2. Solid phase solid phase and char solid phase). The gas phase was assumed to include
The conservation equations for the mass of the solids phase s could
eight components only: CO, CO2, H2, CH4, O2, H2O, tar and N2. Since the
be written as: temperature of introduced wood was higher than the wood drying
  X temperature, the drying process was not modeled in this project. For
∂εs ρs ! ns
Xn  
þ ∇  ε s ρs v s ¼ msl − s¼1
s
mi;s : ð6Þ simplification, the dried woods were assumed to enter the gasifier
∂t l¼1
with 10% wet basis of moisture in the form of water vapor at the fuel
inlet.
The solid phase s momentum equation could be written as: Modeling of pyrolysis of wood was the most challenging step in the
model. Since pyrolysis was a very complicated thermochemical process,
∂ !
 
!!

two one-step global reactions were applied to model both primary and
ε s ρs v s þ ∇  ε s ρs v s v s ð7Þ
∂t secondary pyrolysis. For simplification, the chemical formulae of char
Xn ! ! !
¼ −∇p þ ∇τs − m≠s s
I sl þ I sg þ εs ρs g : and all non-condensable hydrocarbon gases were considered as pure
carbon and methane respectively. The composition of tar was usually
relevant to condensed aromatics, so it was quite reasonable to model
The energy equation in solid phase s could be expressed as:
tar as phenol [31]. Since the temperature was high enough, it was as-
sumed that all tars that produce from the primary pyrolysis were
∂εs ρs H s
þ ∇  ðεs ρs vs H s Þ cracked in the secondary pyrolysis [22]. Due to the complex composi-
∂t Xn h   i  tion of the products of both primary and secondary pyrolysis, the
¼ −∇qs − s¼1 s
hgs T s −T g þ msg hsg −ΔH s : ð8Þ
composition of the product from pyrolysis of wood was developed
from the experimental data available in the literature. Table 1 showed
Since the optical thickness of the gasifier was large, P-1 model the product's mass fraction for both primary and secondary pyrolysis
worked reasonably well to compute the radiative heat transfer [29]. from experimental data.
The P-1 model assumed the propagation of radiation energy into an After the mass balance of the chemical components, the stoichio-
orthogonal series of spherical harmonics [30]. metric coefficients of each chemical component in the two one-step
90 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

Table 1 Table 3
Mass fractions of gas species for primary and secondary pyrolysis from experimental data Reaction rates and kinetic parameters for the char heterogeneous reactions.
[24].  
(5) r 5 ¼ 1:33T ; exp − 1:4710
8
½H2 O0:6 [35]
RT
CO CH4 CO2 H2O H2  
(6) 1:62108
r 6 ¼ 4:4T ; exp − RT ½CO2 0:6 [35]
Primary pyrolysis 0.270 0.056 0.386 0.256 0.032
(7) Kr Kd [24]
Secondary pyrolysis 0.5633 0.0884 0.1110 – 0.0173 r7 ¼ ½O2 
K r þK d  7

ShDg
K r ¼ 2:3T ; exp − 9:2310
RT ; Kd ¼ dp
1:75
Dg ¼ 3:13ðT s =1500
 Þ p0 =p; Sh ¼ 2 þ 0:6Re Sc
1=2 1=3

Re ¼ dp ρg vg −vs =μ g


pyrolysis reactions were determined. The chemical equations of both
primary and secondary pyrolysis were shown as follows:

dried wood→2char þ 0:39CO2 þ 0:42CO þ 0:7H2 þ 0:22CH4 that was surrounded by a stagnant boundary layer, so the gas species
þ 0:15H2 O þ 0:84tar ð10Þ must diffuse through the layer before the char surface reactions occur
[33]. The overall rate of partial oxidation of char with O2 (Reaction 7)
tar→0:30CO2 þ 2:43CO þ H2 þ 0:67CH4 : ð11Þ was expressed as the mixture of kinetic and mass transfer diffusion
controlled [24]. A User Defined Function written in C programming
language was used to incorporate the sub-model. On the other hand,
Arrhenius reaction rates for both primary and secondary pyrolysis
the kinetic rates of char heterogeneous reactions with CO2 and H2O
were [24,32]:
(Reactions 5 and 6) were much slower than the partial oxidation of
! char, so these two reactions were assumed to be kinetically controlled
6 7:78  107 reactions only. The rate expression for char heterogeneous reactions
r p1 ¼ 6:48  10 exp − ½wood ð12Þ
RT was shown in Table 3.

 
6 551218 C þ CO2 →2CObReaction5N ð18Þ
r p2 ¼ 3:08  10 exp − ½tar: ð13Þ
RT

C þ H2 O→CO þ H2 bReaction6N ð19Þ


2.3.2. Homogeneous reaction
Since the flow of gas phase was turbulent, homogeneous gas phase
reactions were controlled by the smaller one of Arrhenius rate and C þ 1=2O2 →CObReaction7N: ð20Þ
eddy-dissipation rate [33]. In this model, four homogeneous gas phase
reactions were introduced, which were the exothermic oxidation reac-
tion of combustible gas: CO, H2 and CH4 (Reactions 1–3) as well as the 2.4. Biomass characteristics
water–gas shift reaction (Reaction 4).
Furniture wood waste Dalbergia Sisoo was used as the biomass
CO þ 1=2O2 →CO2 bReaction1N ð14Þ in the gasification simulations. The proximate analysis and ultimate
analysis of the Dalbergia Sisoo were reported in Table 4. Based on the
ultimate analysis of the wood, the wood was modeled as the hydrocar-
H2 þ 1=2O2 →H2 ObReaction2N ð15Þ bon CH1.53O0.69.

CH4 þ 1=2O2 →CO þ 2H2 bReaction3N ð16Þ 2.5. Solution methodology

The commercially available software FLUENT 14.0 was used to


CO þ H2 O↔H2 þ CO2 bReaction4N: ð17Þ
simulate the biomass gasification. To avoid instability and poor conver-
gence, an unsteady model was used with a small time step size of
The water–gas shift reaction was modeled as a chemical equilibrium 1.0−6 s. It took around 2.5–3 s for this gasification to reach steady
reaction that favored forward reaction at relatively low temperature to state, so for each case the physical time around 7 s was simulated. The
produce more H2 and CO2 and reversed reaction at high temperature to phase-coupled SIMPLE algorithm was used for the pressure–velocity
produce more CO and H2O [23]. The Arrhenius rate expression for coupling, and the second-order upwind scheme was adopted to
homogeneous gas phase reactions was shown in Table 2. discretize the governing equations, and a residual of less than 10−4 for
all the variables was imposed as the stopping criterion. For simplicity,
2.3.3. Heterogeneous reaction the bubbling fluidized bed was considered as a two-dimensional geom-
The heterogeneous reactions between char and gases (O2, H2O and etry, which was discretized with 4390 cells with maximum mesh
CO2) were modeled. Char particle was assumed as a spherical particle interval of 0.005 m, as shown in Fig. 1.

Table 2 Table 4
Reaction rates and kinetic parameters for the homogeneous gas phase reactions. Characteristics of the wood of Dalbergia Sisoo [36].

  Fixed Carbon (FC) Volatile Matter (VM) Ash Calculated LHV (MJ/kg)
(1) r 1 ¼ 3:98  1014 ; exp − 1:6710
8
½CO½O2 0:25 ½H2 O0:5 [5]
RT
  Proximate analysis (% by wt. dry basis)
(2) 13
r 2 ¼ 2:196  10 ; exp − RT 1:09108
½H2 ½O2  [40]
15.70 80.40 3.90 17.83
 
(3) r 3 ¼ 4:40  1011 ; exp − 1:2610
7
½ 0:5
½O2 1:25 [5]
RT CH 4
    Carbon Hydrogen Oxygen Nitrogen
(4) r 4 ¼ 2:78  106 ; exp − 1:2610
7
½CO½H2 O− ½CO 2 ½H2  [22]
RT K p ðT Þ
 7
 Ultimate analysis (% by wt. dry basis)
K p ðT Þ ¼ 0:0265 ; exp 4:54610
RT 48.6 6.2 44.87 0.33
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 91

3. Results and discussion

3.1. Model validation

As highlighted by Gomez-Barea [11], in the field of biomass gasification, the solid validation for numerical simulation was often lacking. In our
study, the gasification was numerically studied inside the novel reactor, which had been experimentally studied at the Institute of Energy Engineering
at Berlin University of Technology [24,34]. But in their experiment, they only focused on the biomass gasification with pure air, so there were no any
data available on the gasification with CO2 as gasifying agent. In order to validate our numerical results with co-gasification of biomass with air and CO2,
the biomass gasification with air as gasifying agent was numerically simulated as the baseline case, and the results were compared with their experi-
mental data, as shown in Fig. 2. It was found that the predicted gas compositions were all slightly underestimated. The deviation of predicted mole
fraction of CO was the maximum, which was about 25%, while the deviation for hydrocarbon gases was the least. The temperature profiles along the
gasifier in both numerical simulation and experiments were also provided, and it was found that the variation trend was quite similar, i.e. with the
increasing height of gasifier the temperature decreases, with largest variations in the middle region of gasifier, this was because the endothermic
gasification reaction mainly happened in this region. The predicted temperatures were about 40–50 K lower than the experimental values. This devia-
tion was quite satisfactory for such gasifier with complex processes inside. As the main reactions and kinetics in the biomass gasifier with air or CO2 were
the same, after validation with the experimental data, our numerical model could be applied to the co-gasification of biomass with air and CO2.

3.2. Effect of CO2-to-wood mass ratio

To synergistically combine the exothermic oxidation reactions of char with the endothermic reactions of CO2 with char, mixtures of air and CO2
with different fractions were used for wood gasification. During the simulation, the wood feeding rate was kept constant at 10 kg/h, the mass
percentage of CO2 in gasifying gas varied from 20 to 80%. It was noted that at fixed CO2 percentage, the increasing CO2/biomass mass ratio meant

0.25

Experiment
0.20 Simulation
Mole fraction

0.15

0.10

0.05

0.00
CO CO2 H2 CH4

(a) Gas compositions


1.2

1.0

0.8
Height h,m

0.6
Experiment
Simulation
0.4

0.2

0.0

-0.2
600 700 800 900 1000 1100 1200
Temperature t,K

(b) Temperature profiles


Fig. 2. Comparisons of main producer gas composition and temperature profiles of the baseline case in numerical simulation with the experimental data [34].
92 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

(K)

(a) Gas temperature


(-)

(b) Gas fraction

(m/s)

(c) Gas velocity

Fig. 3. Distributions of gas temperature, fraction and velocity at 60 wt.% CO2 in the gasifying agent and CO2-to-biomass mass ratio 2.056.
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 93

980

960

Outlet gas temperature T, K


940 20% CO2
40% CO2
920 60% CO2
80% CO2

900

880

860
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2 /biomass mass ratio

Fig. 4. Effect of CO2-to-wood mass ratio on the outlet gas temperature under different CO2 percentages in gasifying agent.

that more CO2 was supplied to the gasifier, and the air supplied to the gasifier would also be increased, as well as the oxygen supply due to its fixed
percentage in air.

3.2.1. Distributions of gas temperature, fraction and velocity


Fig. 3 showed the temperature contour plots of gas phase and char solid phase within the gasifier at 60 wt.% CO2 in the gasifying agent and CO2-
to-biomass mass ratio 2.056. It was seen that at the lower region of the fluidized bed, the temperature distribution was quite uniform due to the
exothermic char oxidation reactions, while in the upper freeboard region, the gas temperature was reduced because the endothermic char gasifica-
tion reactions with CO2 and H2O mainly occurred in this region. From Fig. 3(c) it was found that near the inlet of woodchips, the gas velocity was much
higher than that in other regions, because rapid devolatilization happened, and in this process the solid woodchips were converted to gases with large
volume expansion.
The effect of CO2-to-biomass mass ratio on producer gas temperature under different mass percentages of CO2 in the gasifying gases was shown in
Fig. 4. For all mass percentages of CO2 in the gasifying agent, the temperature of producer gas increased with increasing CO2-to-biomass ratio, while
the producer gas temperature decreased with increasing mass percentage of CO2 in gasifying agent at any CO2-to-biomass ratio. With increasing CO2
percentage in the gasifying gas or decreasing CO2-to-wood ratio, less oxygen was available for the exothermic oxidation reactions of char and vola-
tiles to produce heat, resulting in lower temperature of producer gases.

3.2.2. Distributions of mole fractions of species


Figs. 5 and 6 showed the distributions and axial variations of gas species within the non-symmetric fluidized bed gasifier at 60 wt.% CO2 in
the gasifying agent and CO2-to-wood mass ratio 2.056 at steady state. It could be found that at the height of 0.05 m, O2 was completely con-
sumed due to the oxidation reactions of char and volatiles of pyrolysis to generate heat. The highest mole fraction of CO2 was observed at
the location near the wood inlet because of the oxidation of CO. Also, mole fraction of CO increased while that of CO2 dropped significantly
after a height of 0.1 m since the endothermic char gasification with CO2 (Boudouard reaction) and reversed water–gas shift reaction were
dominated. Because the mole fraction of CH4 was mainly influenced by the pyrolysis of wood which mostly occurred in the lower part of
the gasifier, the rise of mole fraction of CH4 was negligible after the height of 0.06 m. On the other hand, mole fraction of H2 slightly
increased after the height of 0.08 m due to the endothermic char gasification with H2O (water–gas shift reaction) to produce H2 and CO. How-
ever, since the rate of reversed water–gas shift reaction was larger than the rate of forward water–gas shift reaction, some of the H2 was con-
sumed with CO2 to produce CO and H2O.

3.2.3. Mole fractions of producer gases


Fig. 7 showed the effect of CO2/biomass mass ratio under different percentages of CO2 in the gasifying agent on the producer gas composition at
the outlet. As the biomass feeding rate was kept at 10 kg/h, the increasing CO2/biomass mass ratio meant that more CO2 was supplied to the gasifier,
as well as more air or oxygen, therefore the reaction temperature was higher, and the endothermic Boudouard reaction (Reaction 5) was favored.
Furthermore, the reverse water–gas shift reaction (Reaction 4) was also enhanced due to high operating temperature. Both reactions could
convert CO2 into CO, so the mole fractions of CO always increased while the trend of CO2 variation was in contrast with the increasing CO2/biomass
mass ratio.
On the other hand, due to the reverse water–gas shift reaction, more CO2 and H2 were consumed to produce CO and H2O at higher CO2-to-biomass
ratio; the mole fractions of H2 showed a decreasing trend. The dependence of mole fraction of CH4 on CO2-to-biomass mass ratio was not significant
because the mole fractions of CH4 were mainly affected by the pyrolysis of wood under constant feeding rate.
It was known that high concentration of CO2 shifted the water–gas shift reaction in the backward direction to produce more CO. Henceforth,
the increased CO2 percentage in the gasifying agent from 20% to 60% was found to increase the production of CO. However, the production of CO
slightly dropped from CO2 composition of 60% to 80% in the gasifying agent. This might be due to excessive CO2, the amount of air introduced into
the gasifier being not enough to maintain the gasifier at a relatively high operating temperature through the oxidation of char and volatiles of
pyrolysis, so the water–gas shift reaction in the backward direction was weakened, leading to low CO production. In addition, the lower operating
temperature for the high percentage of CO2 was less favorable for the endothermic Boudouard reaction, so it also led to the reduction in CO
production.
94 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

Mole Fraction
Mole Fraction

(a) CO
(c) O2
Mole Fraction
Mole Fraction

(b) CO2 (d) H2


Fig. 5. Distributions of mole fractions of gas species at 60 wt.% CO2 in the gasifying agent and CO2-to-biomass mass ratio 2.056.

When CO2 percentage in gasifying agent was lowered from 80% to 40%, mole fractions of H2 increased because the decreased percentage of CO2
relative to air reduced the reverse water–gas shift reaction, but when CO2 percentage was decreased further from 40% to 20%, the mole fractions of H2
decreased because more H2 was consumed through the oxidization reaction. On the other hand, mole fractions of CH4 increased from CO2
composition of 20% to 60% in gasifying agent due to less contribution of oxygen towards oxidation of CH4, and then decreased from CO2 com-
position of 60% to 80% in the gasifying agent since the rate of wood pyrolysis decreased with the decreasing operating temperature. At a given
CO2-to-biomass ratio, mole fractions of CO2 increased with increasing mass percentage of CO2 in the gasifying agent due to more unreacted
CO2.
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 95

Mole Fraction

(e) CH4
Mole Fraction

(f) N2
Fig. 5 (continued).

From the above analysis it was found that the concentrations of CO and CH4 reached the maximum when the CO2 mass percentage was 60% in
the gasifying agent, and the concentrations of H2 reached the maximum when CO2 percentage was 40%. As the sum of CO and CH4 concentrations
was much higher than that of H2, it might be considered that 60% CO2 percentage in the gasifying agent was the optimal operating condition for
this gasifier, which could also be the proof-of-concept design from the view-point of cold gas efficiency and low heating value, as seen in Figs. 8 and 9.

3.2.4. Lower heating value and cold gas efficiency


Lower heating value (LHV) is defined as the amount of heat released by fully combusting a specified quantity without the latent heat of
vaporization of the water in the combustion product [4].
The lower heating value of fuel LHVfuel can be estimated by [37]:

LHVfuel ¼ 33:9Y C þ 102:9Y H −11:2Y O −2:5Y H2 O ½MJ=kg ð21Þ

where YC, YH, YO, and Y H2 O are the mass fractions of carbon, hydrogen, oxygen and water in the dried biomass.
96 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

0.8
CO
0.7 CO2

Mole fractions of gas species


H2
0.6
CH4
0.5 O2
N2
0.4

0.3

0.2

0.1

0.0
0.0 0.2 0.4 0.6 0.8 1.0
Axial locations L,m

Fig. 6. Axial distributions of mole fractions of gas species at 60 wt.% CO2 in the gasifying agent and CO2-to-wood mass ratio 2.056.

The lower heating value of producer gas, LHVgas, can be calculated as [37]:
h i
3
LHVgas ¼ 10:8yH2 þ 12:6yCO þ 35:8yCH4 MJ=m ð22Þ

where yH2 ; yCO ; and yCH4 are the mole fractions of hydrogen, carbon monoxide and methane in the producer gas.

0.20
0.07
20% CO2
40% CO2
0.15 60% CO2
0.06
CO mole fraction

80% CO2
H2 mole fraction

0.10 0.05
20% CO2
40% CO2
0.05 60% CO2 0.04
80% CO2

0.00 0.03
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2 /biomass mass ratio CO2/biomass mass ratio

(a) CO (c) H2

0.50 0.10
20% CO2 20% CO2
40% CO2 40% CO2
0.45 0.08
60% CO2 60% CO2
80% CO2
CH4 mole fraction

80% CO2
CO2 mole fraction

0.40
0.06

0.35
0.04
0.30
0.02
0.25

0.00
0.20 1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2/biomass mass ratio
CO2 /biomass mass ratio

(b) CO2 (d) CH4

Fig. 7. Effect of CO2/biomass mass ratio on the producer gases under different CO2 percentages in gasifying agent.
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 97

0.70

0.65
20% CO2
0.60 40% CO2

Cold gas efficiency


60% CO2
0.55 80% CO2

0.50

0.45

0.40

0.35

0.30
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2 /biomass mass ratio

Fig. 8. Effect of CO2/biomass ratio on the cold gas efficiency under different CO2 percentages in gasifying agent.

Cold-gas efficiency (CGE) is the ratio of heating value in the cooled producer gas to the energy content of biomass [38], which does not include the
sensible heat of the hot producer gas. It can be calculated as:

V gas LHVgas
CGE ¼  100% ð23Þ
mfuel LHVfuel

where

Vgas volumetric flow rate of producer gas, m3/s


mfuel mass flow rate of fuel, kg/s
LHVgas lower heating value of producer gas, MJ/m3
LHVfuel lower heating value of fuel, MJ/kg.

Figs. 8 and 9 showed the effect of gasifying agent flow rate on LHVgas and CGE, respectively under different mass percentages of CO2 in the gas-
ifying agent. It could be found that the predicted LHVgas and CGE increased with increasing CO2-to-biomass mass ratio for all mass percentages of CO2
in gasifying agent. LHVgas was calculated using the mole fraction of CO, H2 and CH4 in the producer gas. Because higher operating temperature could
be achieved with higher CO2-to-biomass ratio, more non-combustible components (CO2, H2O) could be converted into combustible components (CO,
H2), the LHVgas increased with increasing CO2-to-biomass mass ratio [38].
According to Eq. (23) CGE was proportional to the product of LHVgas and producer gas flow rate, as both the producer gas flow rate and LHVgas
increased with CO2-to-biomass ratio, the increase of CGE with CO2-to-biomass ratio was more significant compared to that of LHVgas. At a given
CO2-to-biomass ratio, LHVgas and CGE increased from CO2 composition of 20% to 60% in the gasifying agent, but they would drop when CO2 percent-
ages increased further to 80% in the gasifying agent. It followed a similar trend to the variation of mole fraction of CO and CH4 with CO2 percentage in
the gasifying agent, because CO and CH4 accounted for a large fraction in the producer gases and were responsible for the higher LHVgas and CGE. As a
result, 60% of CO2 in the gasifying agent could be defined as the optimal percentage, where CO2 addition positively compensated for the burnt fuel gas,
due to its highest LHVgas and CGE.

5
3
Low heating value,MJ/m

20% CO2
2 40% CO2
60% CO2
80% CO2
1
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2 /biomass mass ratio

Fig. 9. Effect of CO2/biomass ratio on the low heating value under different CO2 percentages in gasifying agent.
98 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

0.6

0.5

CO2 conversion ratio


0.4

0.3
20% CO2
40% CO2
0.2
60% CO2
80% CO2
0.1
1.9 2.0 2.1 2.2 2.3 2.4 2.5 2.6
CO2 /biomass mass ratio

Fig. 10. Effect of CO2/biomass ratio on the CO2 conversion ratio under different CO2 percentages in gasifying agent.

3.2.5. CO2 conversion ratio


CO2 conversion ratio is the ratio of the amount of CO2 converted in the gasification process to the amount of CO2 introduced into the gasifier [19].

moles of CO2 in gasifying agent−moles of CO2 in producer gas


X CO2 ¼
moles of CO2 in gasifying agent
ð24Þ
moles of CO2 in gasifying agent−moles of producer gas  mole fraction of CO2
¼ :
moles of CO2 in gasifying agent

The effect of CO2-to-biomass mass ratio on fractional CO2 conversion ratio for different mass percentages of CO2 in the gasifying agent was shown
in Fig. 10. Because larger CO2-to-biomass mass ratio could lead to higher operating temperature, which contributed to higher rate of the endothermic
Boudouard reaction and reversed water–gas shift reaction, the predicted CO2 conversion ratios increased with increasing CO2-to-biomass ratio for all
mass percentages of CO2 in the gasifying agent. At a given CO2-to-biomass ratio, CO2 conversion ratios increased with increasing mass percentages
of CO2 in the gasifying agent because the higher concentration of CO2 favored the Boudouard reaction and reverse water–gas shift reaction, so the
utilization of CO2 for char gasification and reverse water–gas shift reaction were relatively higher [19].

3.3. Effect of wood moisture contents

The moisture content of biomass was one of the main characteristics which affected gasification performance greatly [39]. From the numer-
ical results in Section 3.2, it was shown that the gasifying agent with 60% of CO2 could produce optimal results for the producer gas composi-
tions, LHVgas as well as CGE. Therefore, in the study on the effect of moisture contents on the gasification process, the gasifying agent was
composed of 40% of air and 60% of CO2 and CO2-to-biomass mass ratio was 2.49, and the moisture content of woodchips ranged from 5% to
20% wet basis.

0.4

0.3 CO
CO2
Mole fraction

H2
CH4
0.2

0.1

0.0
0 5 10 15 20 25
Moisture contents, %

Fig. 11. Effect of moisture contents in the woodchips on the mole fractions of producer gases.
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 99

0.8 8
Cold gas efficiency
CO2 conversion rate

3
Low heating value (LHV),MJ/m
CGE/CO2 conversion ratio
Low heating value 6
0.7

0.6
2

0.5 0
5 10 15 20
Moisture contents, %

Fig. 12. Effect of moisture contents in the woodchips on the cold gas efficiency, CO2 conversion ratio and low heating value.

3.3.1. Mole fractions of producer gases


Fig. 11 showed the effect of wood moisture content on the producer gas composition. It could be observed that the mole fraction of H2 increased
while that of CO decreased with moisture content varying from 5% to 20% wet basis. However, with the same variation in moisture content, the in-
crease in the mole fraction of CO was found to be more pronounced as compared to that in the mole fraction of H2 with the same variation interval in
moisture content [40]. As expected, mole fraction of CO2 also increased with moisture content from 5% to 20% wet basis. This was because the increase
of partial pressure of H2O vapor led to an increased rate of the forward water–gas shift reaction and a decreased rate of the reversed water–gas shift
reaction, i.e. more CO2 and H2 was produced than CO and H2O. Also, since the operating temperature decreased with moisture content, higher mois-
ture content was less favorable for the endothermic char gasification with CO2 and H2O to produce CO. Moreover, mole fraction of CH4 decreased with
moisture content from 5% to 20% wet basis because more energy was lost as sensible heat of water vapor, thus it weakened the pyrolysis of wood,
leading to lower CH4 production.

3.3.2. Lower heating value, cold gas efficiency and CO2 conversion ratio
Besides the compositions of producer gas, the moisture content also affected the LHVgas, CGE and CO2 conversion ratio greatly. The effects of wood
moisture content on LHVgas, CGE and CO2 conversion ratio were shown in Fig. 12. The predicted LHVgas decreased with increasing moisture content of
wood from 5% to 20% wet basis, because there was a greater reduction in mole fraction of CO than the increase in mole fraction of H2 [41]. Since the
producer gas flow rate increased while LHVgas decreased with moisture content, the decrease of CGE with moisture content was less significant as
compared to LHVgas [42]. At a constant percentage of CO2 in the gasifying agent, the numerical results showed that mole fraction of CO2 increased
with increasing wood moisture content, but the CO2 conversion ratio showed a decreasing trend. Therefore, in order to achieve high fractional
CO2 conversion, it was still recommended to dry the high moisture wood to a certain degree before introducing the wood into the gasifier.

3.4. Effect of woodchip size

The particle size of wood was also an important parameter which affected the quality of producer gas and gasification performance. The
preparation and reduction of biomass to a required size were associated with an energy cost [43]. The effect of particle size of wood was

0.6
CO
0.5 CO2
H2
0.4 CH4
Mole fraction

0.3

0.2

0.1

0.0
0 2 4 6 8 10
Wood particle size,mm

Fig. 13. Effect of wood particle size on the mole fractions of producer gases.
100 Y. Cheng et al. / Powder Technology 296 (2016) 87–101

1.0 8
Cold gas efficiency
0.9 CO2 conversion rate

CGE/CO2 conversion ratio

3
Low heating value

Low heating value, MJ/m


6
0.8

0.7 4

0.6
2
0.5

0.4 0
0 2 4 6 8 10
Wood particle size, mm

Fig. 14. Effect of wood particle size on the cold gas efficiency, CO2 conversion ratio and low heating value.

investigated with moisture content of 10% wet basis and 0.28 m/s inlet velocity of gasifying agent (40% of air and 60% of CO2) on gasification
performance in the fluidized bed gasifier. For convenience, it was assumed that the particle size of chars, which were produced from the py-
rolysis of wood and/or used as bed materials, was half of the particle size of woods [24].

3.4.1. Mole fractions of producer gases


The effect of wood particle size on producer gas composition was shown in Fig. 13. It was found that the mole fraction of CO2 increased
while those of CO, H2 and CH4 decreased with increasing wood particle size from 2 to 8 mm. Because partial oxidation of char with O2 was as-
sumed to be controlled by both kinetics and mass transfer, an increase in char particle size inhibited the mass transfer of O2 to the char surface.
This led to more O2 being consumed by the oxidation reactions of volatiles to increase the production of CO2. On the other hand, with larger
wood particles, the overall heat exchange surface area was reduced, hence the effective transfer of heat was weakened from gas phase to
both wood and char particle phases, then the wood particles would experience less complete endothermic pyrolysis, resulting in a decrease
of CO, H2 and CH4 production [44]. Also, since the temperature of char phase decreased with increasing char particle size, the rates of endother-
mic char gasification with CO2 and H2O decreased with increasing particle size from 2 to 8 mm. Thus, it led to the reduction in CO and H2 pro-
duction and CO2 conversion.

3.4.2. Lower heating value, cold gas efficiency and CO2 conversion ratio
The effect of wood particle size on LHVgas, CGE and CO2 conversion ratio was shown in Fig. 14. It was found that LHVgas decreased with in-
creasing wood particle size from 2 to 8 mm due to the decreased mole fraction of CO, H2 and CH4, similar to the trend observed by Yin et al. [43].
Furthermore, both LHVgas and CO2 conversion ratio were decreased with increasing wood particle size. Hence it was shown that smaller wood
and char particle sizes were preferred in order to improve the performance of the gasifier.

4. Conclusions effects on the gasification performance in that the lower heating


value of producer gas, cold gas efficiency and CO2 conversion ratio
In our study, the Eulerian method was used to study the CO2 co- was all reduced with increasing moisture contents and particle size.
gasification of biomass inside the fluidized bed gasifier after validation
with the experimental data. The effects of CO2 percentages in the gasify- This study might offer insights on biomass co-gasification with CO2,
ing mixture with air, the CO2-to-biomass ratio, the moisture contents of and aid designs of biomass gasifiers with high performance.
woodchips, and the particle size of woodchip and char on the gasifica-
tion performance were evaluated in terms of the gas compositions Acknowledgment
and temperature, lower heating values, cold gas efficiency and CO2 con-
version ratio. The following observations were made throughout the This research program is funded by the National Research Founda-
study: tion (NRF), Prime Minister's Office, Singapore under its Campus for
Research Excellence and Technological Enterprise (CREATE) program.
1 With increasing CO2-to-biomass mass ratio, more oxygen would be Grant number R-706-001-101-281, National University of Singapore.
available for the exothermic oxidation reactions, resulting in higher
producer gas temperature, and higher mole fractions of CO in the pro- References
ducer gas, but lower mole fractions of H2 and CO2.
[1] C.S. Song, Global challenges and strategies for control, conversion and utilization
2 At 60 wt.% of CO2 in the gasifying agent, the mole fractions of CO and of CO2 for sustainable development involving energy, catalysis, adsorption and
CH4 could reach the maximum. The lower heating value of producer chemical processing, Catal. Today 115 (2006) 2–32.
gas and cold gas efficiency also had the highest values, so this was [2] M.V. Hoeven, CO2 Emissions from Fuel Combustion: Highlights, International Energy
Agency, Paris, 2011.
the optimal operating condition for the biomass gasification with CO2. [3] P.N. Sheth, B.V. Babu, Experimental studies on producer gas generation for wood
3 The moisture content and particle size of woodchips had negative waste in a downdraft biomass gasifier, Bioresour. Technol. 100 (2009) 3127–3133.
Y. Cheng et al. / Powder Technology 296 (2016) 87–101 101

[4] P. Basu, Biomass Gasification and Pyrolysis—Practical Design and Theory, Elsevier, [38] C. Erlich, T.H. Fransson, Downdraft gasification of pellets made of wood, palm-
2010. oil residues respective bagasse: experimental study, Appl. Energy 88 (2011)
[5] S. Murgia, M. Vascellari, G. Cau, Comprehensive CFD model of an air-blown coal- 899–908.
fired updraft gasifier, Fuel 101 (2012) 129–138. [39] T.H. Jayah, L. Aye, R.J. Fuller, D.F. Stewart, Computer simulation of a downdraft wood
[6] S. Kaewluan, S. Pipatmanomai, Gasification of high moisture rubber woodchip with gasifier for tea drying, Biomass Bioenergy 25 (2003) 459–469.
rubber waste in a bubbling fluidized bed, Fuel Process. Technol. 92 (2011) 671–677. [40] I.S. Antonopoulos, A. Karagiannidis, A. Gkouletsos, G. Perkoulidis, Modelling of a
[7] R. Saidur, E.A. Abdelaziz, A. Demirbas, M.S. Hossain, S. Mekhilef, A review on biomass downdraft gasifier fed by agricultural residues, Waste Manag. 32 (2012)
as a fuel for boilers, Renew. Sust. Energ. Rev. 15 (2011) 2262–2289. 710–718.
[8] Z.A.B.Z. Alauddin, P. Lahijani, M. Mohammadi, A.R. Mohamed, Gasification of [41] Z.A. Zainal, R. Ali, C.H. Lean, K.N. Seetharamu, Prediction of performance of a down-
lignocellulosic biomass in fluidized beds for renewable energy development: a draft gasifier using equilibrium modeling for different biomass materials, Energy
review, Renew. Sust. Energ. Rev. 14 (2010) 2852–2862. Convers. Manag. 42 (2001) 1499–1515.
[9] A.V. Bridgewater, Review of fast pyrolysis of biomass and produce upgrading, [42] J.K. Ratnadhariya, S.A. Channiwala, Three zone equilibrium and kinetic free
Biomass Bioenergy 38 (2012) 68–94. modeling of biomass gasifier — a novel approach, Renew. Energy 34 (2009)
[10] G. Berndes, M. Hoogwijk, R. van den Brokek, The contribution of biomass in the future 1050–1058.
global energy supply: a review of 17 studies, Biomass Bioenergy 25 (2003) 1–28. [43] R.Z. Yin, R.H. Liu, J.K. Wu, X.W. Wu, C. Sun, C. Wu, Influence of particle size on perfor-
[11] A. Gomez-Barea, B. Leckner, Modeling of biomass gasification in fluidized bed, Prog. mance of a pilot-scale fixed-bed gasification system, Bioresour. Technol. 119 (2012)
Energy Combust. Sci. 36 (2010) 444–509. 15–21.
[12] C.D. Blasi, Modeling chemical and physical processes of wood and biomass pyrolysis, [44] J.F. Pérez, A. Melgar, P.N. Benjumea, Effect of operating and design parameters on
Prog. Energy Combust. Sci. 34 (2008) 47–90. the gasification/combustion process of waste biomass in fixed bed downdraft
[13] M. Puig-Arnavat, J.C. Bruno, A. Coronas, Review and analysis of biomass gasification reactors: an experimental study, Fuel 96 (2012) 487–496.
models, Renew. Sust. Energ. Rev. 14 (2010) 2841–2851.
[14] H.P. Cui, J.R. Grace, Fluidization of biomass particles: a review of experimental
multiphase flow aspects, Chem. Eng. Sci. 62 (2007) 45–55. Nomenclature
[15] H. Butterman, M.J. Castaldi, CO2 as a carbon neutral fuel source via enhanced
biomass gasification, Environ. Sci. Technol. 43 (2009) 9030–9037. C: linear-anisotropic phase function coefficient
[16] M.F. Irfan, M.R. Usman, K. Kusakabe, Coal gasification in CO2 atmosphere and its Cp: specific heat capacity, J/kg-K
kinetics since 1948: a brief review, Energy 36 (2011) 12–40. d: diameter, m
[17] C. Gonzalo-Tirado, S. Jimenez, J. Ballester, Gasification of a pulverized sub-bituminous Dg: non-dimensional parameter for temperature and pressure corrections
coal in CO2 at atmospheric pressure, Combust. Flame 159 (2012) 385–395. g: gravitational constant, m/s2
[18] C. Gonzalo-Tirado, S. Jimenez, J. Ballester, Kinetics of CO2 gasification for coals of differ- G: incident radiation, W/m2
ent ranks under oxy-combustion conditions, Combust. Flame 160 (2013) 411–416. h: heat transfer coefficient, W/m2-K
[19] T. Renganathat, M.V. Yadav, S. Pushpavanam, R.K. Voolapalli, Y.S. Cho, CO2 utilization H: total enthalpy, J/kg
for gasification of carbonaceous feedstocks: a thermodynamic analysis, Chem. Eng. ΔH: heat of reaction, J/kg
Sci. 83 (2012) 159–170. I: momentum exchange term, kg/m2-s2
[20] T. T. Mani, N. Mahinpey, P. Murugan, Reaction kinetics and mass transfer studies of K: phase exchange coefficient, kg/m3-s
biomass char gasification with CO2, Chem. Eng. Sci. 66 (2011) 36–41. Kd: diffusion rate
[21] L. Garcia, M.L. Salvador, J. Arauzo, R. Bilbao, CO2 as a gasifying agent for gas produc- Kr: kinetic rate
tion from pine sawdust at low temperature using a Ni/Al coprecipitated catalyst, Kp: chemical equilibrium constant
Fuel Process. Technol. 69 (2001) 157–174. m: mass flow rate, kg/s
[22] M. Oevemann, S. Gerber, F. Behrendt, Euler–Lagrange/DEM simulation of ṁ: net production rate/mass transfer rate, kg/m3-s
wood gasification in a bubbling fluidized bed reactor, Particuology 7 (2009) ng: number of gas phase
307–316. ns: number of solid phase
[23] J. Xie, W.Q. Zhong, B.S. Jin, Y.J. Shao, H. Liu, Simulation on gasification of forestry p: pressure, Pa
residues in fluidized beds by Eulerian–Lagrangian approach, Bioresour. Technol. q: heat flux, J/m2-s
121 (2012) 36–46. r: reaction rate
[24] S. Gerber, F. Behrendt, M. Oevermann, An Eulerian modeling approach of wood R: universal gas constant, J/kmol-K
gasification in a bubbling fluidized bed reactor using char as bed material, Fuel 89 Re: Reynolds number
(2010) 2903–2917. SG: user-defined radiation source, W/m3
[25] Z.A. El-Rub, Biomass Char as In-situ Catalyst for Tar Removal in Gasification Sc: Schmidt number
Systems(Ph.D thesis) University of Twente, The Netherlands, 2008. 2008. Sh: Sherwood number
[26] S. Hosokar, K. Kumabe, M. Ohshita, K. Norinaga, C.Z. Li, J.I. Hayashi, Mechanism of t: time, s
decomposition of aromatics over charcoal and necessary condition for maintaining T: temperature, K
its activity, Fuel 87 (2008) 2914–2922. v: velocity, m/s
[27] D. Gidaspow, Multiphase Flow and Fluidization—Continuum and Kinetic Theory V: volumetric flow rate, m3/s
Descriptions, Academic Press, Boston, 1994. X: conversion ratio
[28] M. Syamlal, T.J. O'Brien, Computer simulation of bubbles in a fluidized bed, AIChE y: mole fraction
Symp. Ser. 85 (1989) 22–31. Y: mass fraction
[29] M. Gräbner, S. Ogriseck, B. Meyer, Numerical simulation of coal gasification at circu-
lating fluidised bed conditions, Fuel Process. Technol. 88 (2007) 948–958. Greek letter
[30] Y. Wang, L. Yan, CFD modeling of a fluidized bed sewage sludge gasifier for syngas,
Asia Pac. J. Chem. Eng. 3 (2008) 161–170. α: absorption coefficient, m−1
[31] L. Gerun, M. Paraschiv, R. Lijeu, J. Bellettre, M. Tazerout, B. Gobel, U. Henriksen, ε: volume fraction
Numerical investigation of the partial oxidation in a two-stage downdraft gasifier, μ: shear viscosity, kg/m-s
Fuel 87 (2008) 1383–1393. ρ: density, kg/m3
[32] C. Mandl, I. Obernberger, F. Biedermann, Modelling of an updraft fixed-bed gasifier σs: scattering coefficient, m−1
operated with softwood pellets, Fuel 89 (2010) 3795–3806. τ: viscous stress tensor, kg/m-s2
[33] L. Yu, J. Lu, X.P. Zhang, S.J. Zhang, Numerical simulation of the bubbling fluidized bed
coal gasification by the kinetic theory of granular flow (KTGF), Fuel 86 (2007) Subscript
722–734.
[34] S. Gerber, M. Oevermann, A two dimensional Euler–Lagrangian model of wood d: diffusion
gasification in a charcoal bed — part I: model description and base scenario, Fuel g: gas phase
115 (2013) 385–400. i: species i
[35] M. Kumar, A.F. Ghoniem, Multiphysics simulations of entrained flow gasification. r: reaction
Part II: Constructing and validating the overall model, Energy Fuel 26 (2012) ref: reference
464–479. s: solid phase s
[36] P.N. Sheth, B.V. Babu, Production of hydrogen energy through biomass (waste
wood) gasification, Int. J. Hydrogen Energy 35 (2010) 10803–10810.
[37] C. Gai, Y. Dong, Experimental study on non-woody biomass gasification in a down-
draft gasifier, Int. J. Hydrogen Energy 37 (2012) 4935–4944.

You might also like