You are on page 1of 11

Energy Conversion and Management 201 (2019) 112143

Contents lists available at ScienceDirect

Energy Conversion and Management


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

CFD simulation of a fluidized bed reactor for biomass chemical looping T


gasification with continuous feedstock

Zhenwei Lia,b,1, Hongpeng Xua,b,1, Wenming Yanga,b, , Anqi Zhoua,b, Mingchen Xua,b
a
Sembcorp-NUS Corporate Laboratory, 1 Engineering Drive 2, 117576, Singapore
b
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, 117575, Singapore

A R T I C LE I N FO A B S T R A C T

Keywords: Integrating the gasification process with the chemical looping technology presents a promising route for biomass
Biomass gasification conversion with the objective to obtain high quality syngas without air separation. In this study, the biomass
Chemical looping gasification gasification with iron-based oxygen carrier and continuous feedstock in the bubbling fluidized bed (BFB) fuel
Hydrogen production reactor has been investigated based on the computational fluid dynamics (CFD). The solid phases including fuel
Fluidized bed
and oxygen carriers are modeled based on the pseudo-fluid assumption. The numerical model integrates the
Computational fluid dynamics
multi-fluid model and the chemical reaction models involving the decomposition and gasification of biomass and
the heterogeneous reactions between gases and metal oxides. The predicted time-varying outlet concentrations
of five gas components agree well with the experimental data from the literature. The impacts of the mixing and
segregation behaviors between two solid phases on the gas composition distribution are analyzed. The effects of
operation temperature, fuel feeding rate and steam content on the chemical looping gasification (CLG) perfor-
mance are also investigated. The concentrations of CO and H2 as well as the gas yield and gasification efficiency
increase while the concentrations of hydrocarbons and CO2 decrease with the escalating temperature because of
the facilitation of higher temperature on the endothermic reactions. Raising the feeding rate of biomass leads to
a higher gasification efficiency with more valuable syngas but a lower carbon conversion efficiency due to the
relatively lower OC-fuel ratio. The gasification atmosphere containing 10–50% of steam also brings remarkable
enhancements on the H2 concentration, gas yield and gasification efficiency.

1. Introduction an energy intensive air separation units (ASU) is usually required for
high quality syngas production, which results in high capital cost and
The growing concerns about the fossil fuel shortages and global energy penalty in the gasification system [11,12]. Unlike the conven-
warming crisis have triggered global efforts to increase the use of re- tional gasification technologies, a CLG system usually consists of two
newable energy and reduce the emissions of anthropogenic greenhouse interconnected reactors: a fuel reactor (FR) and an air reactor (AR),
gases. Biomass is regarded as one of the promising alternatives to ad- with the oxygen carriers circulating between them. The recycling of OC
dress these issues thanks to its renewability, abundant availability and can save the cost of oxygen production and avoid the direct contact
carbon neutral characteristic [1,2]. The utilization of biomass can be between fuel and air, thus eliminating both the potential generation of
very extensive and generating valuable syngas through the gasification thermal NOx and the air dilution to syngas [13]. The external re-
process is undoubtedly a favorable energy conversion technology [3,4]. circulation of hematite also functions as the heat carrier which is able to
The gasifying agents providing the oxygen source for biomass gasifi- provide the gasification-required heat in the FR after being heated up in
cation not only can be gaseous agents including air, pure oxygen (O2), the AR [14]. As an innovative technology with such advantages, CLG
steam (H2O), carbon dioxide (CO2) [5–8], but also can be solid particles has gained the interests of more and more researchers and has been
such as metal oxides, which serve as the oxygen carriers (OC) in the tested with various biomass fuels such as microalgae [15], pine sawdust
novel technology named chemical looping gasification (CLG) [9,10]. [16], wheat straw [17], rice straw [18] and rice husk [19], together
Previous researchers indicated that air-blown gasifiers produce low with different oxygen carriers including iron-based materials [15–17],
quality syngas due to the dilution of air, so that high purity oxygen from combined Fe-Cu oxides [18] and manganese ore [19]. Zeng et al. [20]


Corresponding author at: Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, 117575, Singapore.
E-mail address: mpeywm@nus.edu.sg (W. Yang).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.enconman.2019.112143
Received 29 July 2019; Received in revised form 30 September 2019; Accepted 1 October 2019
Available online 10 October 2019
0196-8904/ © 2019 Elsevier Ltd. All rights reserved.
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

Nomenclature t time, s

v velocity, m/s
bi stoichiometric factor for reaction i X conversion of solid
C gas concentration, mol/m3; relative volume fraction Y mass fraction of species
Cd drag coefficient
ds particle diameter, m Greek letters
ess coefficient of restitution
E activation energy, kJ/mol α volume fraction
fc carbon fraction βsg the fluid-solid exchange coefficient
g0,ss radial distribution function γΘm collisional dissipation, kJ/(m3·s)
G gas yield, Nm3/kg ε0 initial porosity of the particle
Ī¯ identity matrix or tensor εs volume fraction of solid
k0 pre-exponential factor, 1/s η efficiency
Keq equilibrium constant Θs granular temperature
ṁ consumption rate, kg/(m3·s) λ bulk viscosity, Pa·s
m mass, kg μ shear viscosity, Pa·s
MW molecular weight, kg/kmol ρm molar density, mol/m3
n reaction order ρ density, kg/m3
Nu Nusselt number τ̄¯ stress tensor
P pressure, Pa Φls energy loss to the fluid, kJ/(m3·s)
Pr Prandtl number
r reaction rate, m/s Subscripts
ro grain radius, m
→ g gas
R momentum transfer, m/s
R universal gas constant, J/(mol·K) i, j species or component
Re Reynolds number p, q phases
Ro oxygen carrying capacity ox oxidized form
S0 initial surface area, m2/m3 red reduced form
T temperature, K s solid

investigated the biomass self-moisture CLG process in a fixed bed re- agreement towards experimental measurements, for the pressure in the
actor and found a noticeable increase in the gas yield owing to the FR as well as the gas composition and temperature in the AR. Mean-
moisture content. Huang et al. [21] carried out the contrast experi- while, several numerical investigations regarding the biomass gasifi-
ments: biomass pyrolysis with quartz sand as bed materials and biomass cation have been reported recently. Luo et al. [28] developed a CFD
gasification in the presence of natural hematite in a bubbling fluidized model to simulate a dual fluidized bed system for biomass gasification,
bed (BFB) fuel reactor. Results showed that the oxygen source provided in which they carried out a comparison of a hybrid EMMS drag model
by the oxygen carriers can convert more carbon into syngas and the and the Gidaspow drag model. Yang et al. [29,30] incorporated the
decrease of tar content also suggested the catalysis of oxygen carriers on CFD-DEM coupling model with heat transfer and chemical reactions to
tar cracking. Wei et al. [22] performed the CLG process using pine study the particle-scale behaviors and explore the effects of bubble
sawdust as fuel and synthesized Fe2O3/Al2O3 as oxygen carrier in a dynamics for the biomass gasification process in the fluidized bed.
10 kWth interconnected circulating fluidized bed reactor. It was de- Ostermeier et al. [31] performed the coarse-grained CFD-DEM simula-
monstrated that the concentrations of CO, H2 and CH4 increased with tion of biomass gasification to investigate the evolution of the wood
the increasing of temperature and the optimal biomass feeding rate was pellet with size of 6 mm and obtained reasonable prediction of the
found to obtain the highest cold gas efficiency. The continuous opera- gasifier behavior and performance. The significance and effectiveness
tion of biomass CLG was further scaled up to a 25 kWth prototype of CFD in the study of multiphase chemical looping systems as well as
composed of a high velocity fluidized bed as an air reactor and a BFB as the biomass gasification process have also attracted increasing attention
the fuel reactor as reported in Ge et al. [23]. A significant improvement among the research and exploration of the CLG process. Wang et al.
of carbon conversion efficiency was observed with the temperature in [32] implemented a CFD simulation of coal char gasification in a fuel
the range of 800–900 °C and the optimal steam-to-biomass ratio was reactor in order to investigate the influence of operating parameters on
found at 1.0 for higher gas yield while maintaining a high efficiency. CLG performance and gas products. It was concluded that a smaller OC
Larsson et al. [24] utilized 12% ilmenite as the catalyst mixing with particle size and a lower operating velocity would reduce the syngas
silica sand in a 2–4 MWth dual fluidized bed (DFB) gasifier and achieved production during the CLG process. Li et al. [33] employed the Eulerian
~50% decrease in the yield of tar, accompanied with the undesirable multiphase approach to simulate the biomass CLG process, where the
reduction in the cold gas efficiency and the heating value of the gas thermal degradation of microalgae played an essential role in the fixed
products. It was also emphasized that the impact of adding ilmenite bed (FB) fuel reactor. With the consideration of the two-step pyrolysis
showed a high dependence upon the operating conditions of the DFB and iron-steam reactions, the model presented in [33] was well vali-
gasifier. Therefore, more efforts are required for the industrial scale and dated by comparing the predicted time-varying concentrations of var-
commercialization of the biomass CLG technology [13]. ious gas species with the experimental data of Liu et al. [15]. None-
The computational fluid dynamics (CFD), as a useful approach to theless, the BFB reactor with the key advantages of excellent heat and
analyze the interaction between hydrodynamics and chemical kinetics, mass transfer, high degree of mixing and ease of scale-up is widely
has been extensively employed to provide detailed information for the applied for chemical looping gasification [34].
operational optimization and scale-up of the chemical looping com- Although the continuous operation of biomass gasification through
bustion (CLC) system [25,26]. May et al. [27] developed CFD models chemical looping is of great importance for its development and com-
for fuel and air reactor of the 1 MWth CLC pilot plant and obtained good mercialization, few reports on the numerical investigations of CLG

2
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

process with continuous feedstock of biomass have been previously expressed as


presented. In this study, a numerical simulation is conducted to in-

vestigate the CLG performance with the sawdust of pine being con- (αs ρs →
vs ) + ∇ ·(αs ρs →→
vs vs )
∂t
secutively fed into the BFB fuel reactor and reacting with the iron-based n
oxygen carriers. The predicted results of the Eulerian multiphase model = −αs ∇p + ∇ ·τ¯s + αs ρs→
g + ∑ (→
Rrs + ṁ rs →
vrs − ṁ sr →
vsr )
integrated with the chemical kinetics including pyrolysis, gasification r=1 (5)
and metal oxide reduction are compared with the measured values from
The stress tensor for the granular solid phase can be obtained by
the experimental study [21]. The influences of the particle behaviors
such as mixing and segregation between different phases in the CLG τ¯s = −ps I¯ + μs αs (∇→
vs + ∇→
v s ) + αs λs (∇ ·→
T
vs ) I¯ (6)
system are analyzed. The validated model is then applied to further
evaluate the effects of operation temperature, biomass feeding rate and where ps represents the solid pressure, μs is the granular viscosity.
steam content on the biomass CLG performance. The fluid-solid momentum exchange term is the opposite number of
the solid-fluid momentum, βgs = −βsg . The Gidaspow drag model [36]
2. Methodology as a combination of the Wen and Yu model [37] and the Ergun equation
[38] is adopted in this work. The drag coefficient is given as
2.1. Details of the experimental study → →
⎧150 αs (1 − αg ) μg + 1.75 ρg αs | vs − vg | , αg ≤ 0.8,
⎪ αg ds2 ds
The experimental data to be used for simulation are provided by βsg =
⎨ 3 αs αg ρg | →
vs − →
vg | −2.65
Huang et al. [21], in which the continuous feedstock of biomass went ⎪ C αg , αg > 0.8
⎩ 4 d ds (7)
through the chemical looping gasification process in the presence of
oxygen carrier. The BFB reactor has a length of 1000 mm and an inner where
diameter of 60 mm. A porous quartz plate loading the metal oxide
particles is placed at the height of 300 mm, thus the simulated reaction 24
Cd = [1 + 0.15(αg Re)0.687]
zone is considered to be in height of 700 mm. The reactor was heated to αg Re (8)
the desired operating temperature, after which the pine sawdust was
The solid stress accounting for the collision between particles is
continuously fed from the top and entered the BFB reactor through a
given as [39]:
drop tube. The oxygen carrier employed was natural hematite with
90 wt% Fe2O3 as the reactive component. Table 1 summarizes the de- ps = αs ρs Θs + 2ρs (1 + ess ) αs2 g0, ss Θs (9)
tailed physical properties and operating parameters of the CLG ex-
perimental study. The ultimate and proximate analyses of the pine where g0, ss represents the radial distribution function at contact and is
sawdust (dry basis) are listed in Table 2. expressed as

g0, ss = [1 − (αs / αs,max )1/3]−1 (10)


2.2. The CFD model
The solid viscosity is given as
The Eulerian approach has been applied to describe the gas and
solid phases in the fluidized bed reactor. The present work is performed μs
based on the Eulerian multiphase model. Three phases were taken into 4 Θ 1/2 10ρs ds Θs π
= αs ρs ds g0, ss (1 + ess ) ⎛ s ⎞ +
consideration in this model, including two solid phases for biomass and 5 π
⎝ ⎠ 96αs (1 + ess ) g0, ss
oxygen carriers and one gas phase. The governing equations describing 2
4
mass, momentum, energy and species transport are solved for each ⎡1 + g0, ss αs (1 + ess ) ⎤
⎣ 5 ⎦ (11)
phase as summarized below [35].
The continuity equation for phase q is written as and the bulk viscosity of solids is obtained by
n
∂ Θ 1/2
(αq ρq ) + ∇ ·(αq ρq→
vq ) = ∑ (ṁ pq − ṁ qp) λs =
4
αs ρs ds g0, ss (1 + ess ) ⎛ s ⎞
∂t p=1 (1) 3 ⎝π⎠ (12)

where αq represents the volume fraction of the qth phase, ṁ pq is the An equilibrium between the production and dissipation of the
mass transfer rate from the pth phase to the qth phase. random kinetic energy is assumed for the calculation of the granular
The transport equation for each species is given as temperature, which is

τs: (∇→
n m
∂ us ) = γΘm + Φls
(αq ρq Yiq) + ∇ ·(αq ρq→
vq Yiq) = ∑ ∑ (ṁ ijqp − ṁ jipq) (13)
∂t p=1 j=1 (2)
where the collisional dissipation is
where Yiq is the mass fraction and i, j denote different species.
The conservation equation of momentum for the gas phase is Table 1
written as Simulated fuel reactor properties [21].

∂ Description Value
(αg ρg →
vg ) + ∇ ·(αg ρg →→
vg vg )
∂t Height of the computational domain of the reactor 700 mm
n
→ Inner diameter of the reactor 60 mm
= −αg ∇p + ∇·τ¯g + αg ρg→
g + ∑ (Rsg + ṁ sg →
vsg − ṁ gs →
vgs ) Inner diameter of the drop tube 20 mm
s=1 (3) Flow rate of Ar from the bottom of the reactor 1500 L/h
→ Mass flow rate of biomass from the drop tube
where Rsg = βsg (→
vs − →
0.12 kg/h
vg ) is the drag term. Mean diameter of sawdust 0.34 mm
The fluid stress tensor is calculated as Mass of the oxygen carrier 150 g
Mean diameter of oxygen carrier 0.215 mm
τ¯g = αg μg (∇→
vg + ∇→
v g ) + αg λ g (∇ ·→
T
vg ) I¯ (4) Temperature maintained by the furnace 840 °C
Operation pressure 1 atm
The conservation equation of momentum for the sth solid phase is

3
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

Table 2
Ultimate and proximate analysis of pine sawdust (dry basis) [21].
Ultimate analysis (wt%) Proximate analysis (wt%) LHV (MJ/kg)

C H O N S Volatiles Fixed carbon Ash 19.80


48.44 6.21 45.29 0.05 0.01 84.82 14.58 0.60

12(1 − ess2 ) g0, ss pyrolysis.


γΘm = ρs αs2 Θ3/2
s
ds π (14)
2.3.2. Char gasification
and the dissipation in fluid is
Φls = 3βs Θs (15) C + CO2 → 2CO (R3)

The solid-solid exchange coefficient K s1 s2 between the two solid C + H2O → CO + H2 (R4)
phases is expressed as [40]:
The kinetic model of Everson et al. [44] is adopted for the char
3(1 + ess )(π /2) α s1 α s2 ρs1 ρs2 (d s1 + d s2 )2g0, ss → gasification rate, which is given as
K s1 s2 = | vs1 − →
vs2|
2π (ρs1 d s31 + ρs2 d s32 ) (16) S0
ṁ Char = ρs εs r (1 − X )2/3
1 − ε0 (20)
The conservation equation of energy is written as
n where ε0 is the initial porosity, S0 represents the initial surface area of
∂ ∂p
(αq ρq hq) + ∇ ·(αq ρq→
vq hq) = αq + τ¯q: ∇ ·→
vq − ∇ ·→
qq + Sq + ∑ (Qpq) char, and r is written as
∂t ∂t p=1
ki Ki Pi
ri =
(17) 1 + Ki Pi + Kj Pj (21)
where hq represents the specific enthalpy of the qth phase, Sq denotes
where i represents the reactant CO2 or H2O, j represents the product CO
the source term for enthalpies because of the chemical reactions, →
qq is or H2, respectively.
the heat flux, Qpq represents the heat transfer between gas and solid
phases: 2.3.3. Water-gas-shift (WGS) reaction
Qpq = hpq Ai (Tp − Tq) (18)
CO + H2O → CO2 + H2 (R5)
where hpq (=hqp ) is the volumetric heat transfer coefficient between the
pth phase and the qth phase and is related to the Nusselt number Nus in The homogeneous WGS reaction is taken into consideration in the
the case of granular flows, which is calculated based on the Gunn CLG process and its reaction rate is written as [45]
correlation [41]:
1 −E / RT
r = −k 0 ⎛⎜e−E / RT CH0.5 C
2 CO2
− e CH2O CCO⎞⎟
Nus ⎝ K eq ⎠ (22)
= (7 − 10αf + 5α f2 )(1 + 0.7Re0.2 1/3 2 0.7
s Pr ) + (1.33 − 2.4αf + 1.2α f )Res 7
where k 0 = 2.17 × 10 1/s, E = 192.9 kJ/mol, and
Pr1/3 (19) K eq = exp( −4.33 + 4577.8/ T ) .

2.3.4. Metal oxide reduction


2.3. Chemical reaction kinetic model

CH4 + 9.75Fe2O3 →
The chemical reactions taking place in the BFB fuel reactor are
0.25CO + 0.5H2 + 0.75CO2 + 1.5H2O + 6.5Fe3O4 (R6)
complicated not only due to the complex thermal degradation process
of biomass which could result in numerous different products [42], but C2H4 + 12Fe2O3 → CO + CO2 + H2 + H2O + 8Fe3O4 (R7)
also because of the high uncertainty to determine the products gener-
ated from the partial oxidation of the fuel by the oxygen carriers [1]. In CO + 3Fe2O3 → CO2 + 2Fe3O4 (R8)
the present study, the following reaction mechanisms were introduced H2 + 3Fe2O3 → H2O + 2Fe3O4 (R9)
to investigate the biomass CLG in the BFB fuel reactor.
The partial oxidation of hydrocarbons (CH4 and C2H4) into CO and
2.3.1. Pyrolysis H2 in the biomass CLG process is suggested by both the experimental
A two-step model considering the biomass devolatilization and tar
cracking was used for the biomass pyrolysis [33,42]. Table 3
Mass fraction of the products from biomass pyrolysis.
Biomass → Char + Woodgas(CH4,C2H4,CO,CO2,H2,H2O) + Tar (R1)
Component Mass fraction
Tar → Woodgas(CH4,C2H4,CO,CO2,H2,H2O) + Tarinert (R2)
Primary pyrolysis Secondary pyrolysis
The mass fraction of the products from the primary pyrolysis (R1)
Char 0.147 –
and secondary pyrolysis (R2) are determined by the proximate and CH4 0.078 0.102
ultimate analyses of the pine sawdust [21] and are listed in Table 3. It is C2H4 0.037 –
assumed that the tar generated from biomass degradation would further CO 0.456 0.650
decompose into syngas and inert tar [43]. The reaction rates of both CO2 0.183 0.128
H2 0.016 0.020
pyrolysis reactions are expressed in the form of Arrhenius law, i.e.,
H2O 0.010 –
k 0 exp(−Ea/ RT ) , with pre-exponential factor k 0, pri = 4.13 × 106 1/s and Tar 0.073 –
activation energy Ea, pri = 112.7 kJ/mol for primary pyrolysis, Tarinert – 0.100
k 0, sec = 9.55 × 104 1/s and Ea, sec = 123.3 kJ/mol for secondary

4
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

studies [15] and numerical investigations [33]. It is also noted that the inside the reactor with the initial height of 26 mm. Pure argon is in-
main component of the reduced metal oxides is Fe3O4 in Huang et al. troduced from the bottom of the reactor as fluidizing gas with the flow
[21], thus the further reduction of oxygen carrier into FeO is not taken rate of 1500 L/h. When the bed materials are fluidized, pine sawdust
into account. The OC reduction rates are calculated based on the and balance gas begin to enter the reactor through the drop tube inlet,
shrinking core model (SCM) for the heterogeneous reactions [46]. The with the flow rate of 0.12 kg/h and 200 L/h, respectively. The gas outlet
conversion degree of oxygen carriers is used to describe the progress of is set at the top of the reactor and assumed to be at atmospheric pres-
the reactions [47]. sure. A constant temperature of 840 °C is assumed for the walls of the
mox − m reactor.
X= A two-dimensional axisymmetric mesh is used for the fuel reactor.
mox − mred (23)
Fig. 1a displays the lower portion of the computational domain. The
Differentiating Eq. (21) gives the conversion rate of the metal oxide, mesh sensitivity analysis, similar to the published work [47], was first
which is related to the mass loss of the OC particles. implemented using two different grid sizes. In order to maintain the
dX 1 dm 1 dm numerical stability, the time step size was also reduced by a factor of
= = five for the fine mesh. Fig. 1b shows the comparison of the outlet gas
dt mox − mred dt R o mox dt (24)
concentrations versus time using fine and coarse mesh in the first 5 min.
where R o = (mox − mred )/ mox represents the oxygen carrying capacity. It can be seen that the discrepancy of the results obtained from two sets
The expressions of the consumption rates of the combustible gases of grids is not evident. Therefore, the coarse mesh is chosen in the
by oxygen carriers are summarized in Li et al. [33]. The reaction rates subsequent simulations because of the significantly lower computa-
of H2 and CO are given as tional cost.
ki R o 3MWFe2O3 ⎞
ṁ i = ρ εs ⎜⎛YFe O + YFe3O4 × ⎟ (1 − X )
2/3MW
i
2MWO2 s ⎝ 2 3 2MWFe3O4 ⎠ (25) 3. Results and discussion
3bi k 0, i e−Ei / RT (Ci − Ci, eq)
ki = 3.1. Model validation
ρm ro (26)
where i stands for H2 and CO, bH2 = bCO = 3, k 0,H2 = 2.3 × 10-3 1/s, Fig. 2 shows the evolution of volume fraction and velocity vectors of
k 0,CO = 6.2 × 10-4 1/s, EH2 = 24 kJ/mol, ECO = 20 kJ/mol. oxygen carriers and biomass with time. Since sawdust was continuously
The consumption rate of CH4 is written as fed into the reactor through the drop tube when the desired tempera-
ture was reached and the oxygen carriers were well fluidized, the time
kCH4 R o 12MWFe2O3 ⎞ YCH4 that marks the start of biomass input is considered to be time zero. At
ṁ CH4 = ρ εs ⎜⎛YFe O + YFe3O4 × ⎟ (1 − X ) M
2MWO2 s ⎝ 2 3 8MWFe3O4 ⎠ YCH4, TGA the time t = −1 s, the bed height reaches about 40 mm and the regions
WCH4 of low solid volume fraction can be observed, which demonstrates the
(27)
formation of bubbles and the reasonably good fluidization state. The
−4
where kCH4 = 5.33 × 10 1/s, YCH4, TGA = 0.1. instantaneous contour of biomass at t = 1 s displays the successful entry
The consumption rate of C2H4 can be obtained with the same of biomass into the reactor, while the corresponding velocity vector
coefficients according to the reaction kinetics provided in [47]. shows the upward motion above the inlet of these particles, which is
ascribed to their low solid density and the relatively high velocity of the
2.4. Initial and boundary conditions and grid independence rising gas stream from the bottom of the reactor. The continuous
feedstock of sawdust will experience instantaneous thermal decom-
The oxygen carriers which contain 90 wt% of Fe2O3 are packed position once they enter the reactor, releasing volatile gases and tar at

Fig. 1. Computational domain and mesh sensitivity analysis. (a) Schematic of the BFB reactor and the magnified lower and upper portions of the coarse compu-
tational meshes. (b) Comparison of the outlet gas concentrations versus time using a fine and a coarse mesh.

5
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

Fig. 2. Instantaneous volume fractions and ve-


locity vectors (m/s) of oxygen carriers and bio-
mass. The oxygen carriers were first fluidized by
argon flow from the bottom of the reactor at
1500 L/h. Biomass was continuously fed from
the drop tube into the reactor from the time
t = 0 s, with a feeding rate of 0.12 kg/h and
operating temperature of 840 °C.

such temperature and remaining char. Although a negligible amount of the experimental system behaves more like CLC owing to the extremely
char may be entrained to the freeboard area at a relatively low velocity, high OC-fuel ratio, generating high concentration of CO2 and low
as recognized in the subsequent contours and vectors, the bed surface of concentrations of CH4, C2H4, CO and H2 in the outlet. In the meantime,
the emulsion phase is mainly in the range of 40–60 mm, which matches CO2 and H2O are favorable gasifying agents for the char conversion into
the bed height described in the experimental study [21]. CO and H2, resulting in low char mass in the first 5 min. The lattice
Comparing the phase contours of oxygen carriers and biomass oxygen availability decreases as the mass fraction of Fe2O3 is gradually
(Fig. 2a and b) it can be observed that the motion behaviors of OC and descending, hence the partial oxidation of fuel begins to dominate in
fuel particles in the bed are different and fluctuating, resulting in dis- the reactor, leading to the increase of combustible gases concentrations
tinct and unstable bed surfaces for the two solid phases. The dense and decrease of CO2 concentration. Correspondingly, the consumption
region of oxygen carriers is usually in the lower portion of the reactor, rate of char declines as the generation of gasifying agents decreases, so
while the fuel particles are mainly floating near the inlet of drop tube that the char mass within the reactor climbs to a higher value as the
because of the progressively lower density, making the upper portion reaction proceeds. In the last 15 min, all gas concentrations are basi-
the primary contact area between the metal oxides and the volatile cally stable over time and the mass variations of OC and char within the
gases produced from pyrolysis as well as the CO and H2 generated from reactor are not obvious as the remaining Fe2O3 gradually decreases to
char gasification. Such segregation of two particle phases was also near zero. Therefore, the CLG fuel reactor can be considered to be in a
observed in the numerical studies of Wang et al. [32] and Armstrong quasi-steady state after 30 min of the reaction stage.
et al. [48]. On the other hand, Mahalatkar et al. [47] indicated that the Fig. 5 shows the instantaneous distributions of gas compositions at
insufficient contact between fuel and OC particles leads to partial
combustion, which is undoubtedly a weakness in chemical looping
combustion but could be considered as an advantage for syngas pro-
duction in the CLG process.
Fig. 3 shows a comparison of the calculated gas concentrations of
five gas components and the measured values as a function of time
[21]. The model predictions provide an excellent agreement with the
experimental data. The significant variations of gas concentrations with
reaction time (especially H2 and CO2) in the CLG process are captured
in a reasonable manner. The concentrations of all the combustible gases
(CH4, C2H4, CO and H2) display a slowly increasing trend during the
test, while the CO2 concentration decreases rapidly with time in the
first 20 min, after which the change gradually becomes gentle.
In the fuel reactor of the CLG process, metal oxide reduction reac-
tions as well as char gasification play an important role in the syngas
compositions. The mass variations of different solid particles within the
reactor are shown in Fig. 4, where the evolutions of hematite and
magnetite mass are represented in the form of line while the change in
char mass is expressed in the form of columns. The use of histogram for
time-averaged char mass per 5 min is to eliminate the strong fluctuation
due to its production during the devolatilization of continuous feed-
stock and its consumption through heterogeneous reactions, thus better
illustrating the evolution of char mass over time. Fig. 3. Simulated concentrations of synthesis gas in comparison with the
In the initial stage of the CLG process, hematite as the bed material measured data of Huang et al. [21], with biomass feeding rate of 0.12 kg/h and
undoubtedly far outweigh the progressive feeding of sawdust, so that temperature of 840 °C.

6
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

fluidizing agent.
Fig. 7 shows the compositions and total yield of syngas along with
the carbon conversion efficiency (ηc), gasification efficiency (η), and
lower heating value (LHV, MJ/Nm3) as the function of operation tem-
perature. They are obtained by the following equations [15,21]:

12(CCO + CCO2 + CCH4 + 2CC2H4 ) Gv


ηC = × 100%
22.4fC (28)

LHV = 12.6CCO + 10.8CH2 + 35.9CCH4 + 63.5CC2H4 (29)

LHV × G V
η= × 100%
Qb (30)

where Gv (Nm3/kg) is the total yield of five gas species (i represents


Fig. 4. Mass of oxygen carriers and char in the fuel reactor as a function of time. CH4, C2H4, CO, CO2 and H2), Ci is the mole fraction of the corre-
sponding gas component in the flue gas, fC is the carbon content of
t = 1 min and t = 10 min. As expected, all of the product gas species biomass, Qb (MJ/kg) is the lower heating value of the sawdust.
only occur around and above the biomass inlet compared with the As can be seen in Fig. 7, an obvious increase from 47.55% to 54.27%
corresponding volume fraction contours in Fig. 2, which means that the of CO and decrease from 14.79% to 8.29% of CO2 composition are quite
metal oxide reduction reactions mainly take place in the upper portion remarkable as the temperature increases from 740 °C to 890 °C. This is
of the bed, with the reaction rates displayed in Fig. 6. It can be seen because higher temperature enhances the endothermic Boudouard re-
from Fig. 5 that the highest mole fractions of CH4, C2H4, CO and H2 are action (R3) and restrains the exothermic reaction (R8), facilitating the
in the emulsion region because of the rapid devolatilization of biomass. CO generation and weakening the conversion of CO into CO2 simulta-
Another reason accounting for the peak concentrations of CO and H2 neously. The enhancement of endothermic steam gasification (R4) by
relies on the endothermic char gasification with CO2 and H2O in the the high temperature also leads to a steady increase of H2 concentra-
emulsion phase. The locations of the peak mole fractions of CO2 and tion. However, the concentrations of the hydrocarbons display a slight
H2O, on the other hand, are slightly higher than those of other gas decrease with the increasing temperature. Similar results were pre-
components. This is because the combustible gases are consumed by the sented in the experimental study [49]. As CH4 and C2H4 oxidations by
metal oxides and more CO2 and H2O are produced in addition to the OC are endothermic reactions, they may consume more of these two gas
thermal degradation of biomass. Furthermore, the increase of mole species at higher temperature.
fractions of CH4, C2H4, CO and H2 and decrease of mole fractions of CO2 In addition, it is observed that the increase in temperature results in
and H2O with time can also be observed by comparing the left hand side a decrease of H2/CO ratio, which is consistent with the inference that
with the right hand side of the contour for each gas species. This can be the H2/CO ratio is adjustable within a certain temperature range in
well explained by Fig. 6, which demonstrates that the reaction rates of biomass CLG as stated in [49]. Furthermore, both the gasification ef-
OC reduction are also declining with the reaction proceeding in the ficiency and carbon conversion efficiency as well as the gas yield in-
biomass CLG reactor. crease with the escalating temperature since the endothermic biomass
conversion including pyrolysis and char gasification are significantly
enhanced at higher temperature. The evolution of LHV, on the other
3.2. Effects of operation temperature hand, is indistinctive due to the combined effect of decreasing hydro-
carbon gases and increasing CO composition.
One of the essential operating conditions for biomass CLG is the The evolution of oxygen carrier conversion X and conversion rate
reactor temperature, which can influence the thermal conversion of dX/dt at various operation temperature are shown in Fig. 8. It can be
biomass and the heterogeneous reactions between OC and the com- seen that at each moment, the OC conversion at higher temperature is
bustible gases. Previous researchers have pointed out that a CLG system always above that at lower temperature in the modelling time of
is better to be operated in the temperature range of 650 °C to 1000 °C, 30 min, attributing to the enhancement of OC reduction reactions at
below which the CLG reactions cannot occur, and exceeding which the high temperature. Their differentiation, on the other hand, shows the
oxygen carriers are likely to sinter [22]. It is also reported that the pores same temperature-related results in the first 10 min but then turns into
among the granules might be blocked due to the fusion of gasification the opposite relationship after the intersection, indicating that the
ash at higher temperature, which is unfavorable for maintaining the conversion rate of oxygen carrier firstly increases with the rising tem-
reactivity of the oxygen carriers [15]. Hence, the range of 740–890 °C perature but descends faster due to the rapidly decreasing mass fraction
was selected to investigate the effects of operation temperature on of Fe2O3.
biomass CLG, with the feeding rate of 0.12 kg/h and argon as the

Fig. 5. Instantaneous contours of gas species mole fractions at t = 1 min (left) and t = 10 min (right).

7
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

Fig. 6. Reaction rates (kmol/(m3·s)) of OC reduction at t = 1 min (left) and t = 10 min (right).

attributed to the OC reduction reactions (R6–R9) since the biomass CLG


is a partial oxidization process. When the sawdust feeding rate is rela-
tively small, the fuel is more prone to be fully converted into CO2 and
H2O by the excessive lattice oxygen provided by metal oxides. As the
biomass feeding rate increases, while the inventory of natural hematite
is kept constant, the reactions between the increasing volatiles released
from biomass pyrolysis and the available lattice oxygen provided by
oxygen carriers become more incomplete, resulting in the rise of the
combustible gases. In the meantime, the increasing char content in the
reactor also benefits both the Boudouard (R3) and steam gasification
(R4) reactions, stimulating the generation of CO and H2 as well as the
consumption of CO2.
As seen from Fig. 9, the gasification efficiency and LHV display a
steady upward trend whereas the carbon conversion efficiency declines
slightly as the fuel feeding rate increases. The descent of carbon con-
version can be explained by the decreasing OC-fuel ratio as the lattice
Fig. 7. Effects of operation temperature on biomass CLG performance. oxygen would be less sufficient due to the larger amount of biomass
entering the reactor. That is to say, part of the residual char might not
be fully converted into carbonaceous gas, thus lowering the carbon
conversion efficiency [22]. The variations of gas compositions, espe-
cially the growth of CO and H2 as well as the decline of CO2, lead to an
increase of LHV. Meanwhile, the change in gas yield is insignificant,
although it is reported that increasing the OC-fuel ratio would con-
tribute to an increase of the gas yield in the experimental study with a
fixed bed reactor [20]. This is because the amount of OC inventory can
be considered fairly larger than that of the consecutive feedstock of
biomass in the BFB reactor, meaning that the OC-fuel ratio is kept at a
relatively high value, so that the gas yield could be maintained by
consuming more lattice oxygen from the oxygen carriers. Therefore, the
increasing feeding rate gives a higher gasification efficiency in the
biomass CLG process.
Fig. 10 illustrates the evolution of oxygen carrier conversion X and
conversion rate dX/dt with various biomass feeding rate. It is observed
that the increase of biomass feeding rate significantly promotes the
Fig. 8. Evolution of OC conversion at different operation temperature. conversion of oxygen carrier owing to the enhanced redox reactions

3.3. Effects of biomass feeding rate

Different from the goal of thorough combustion in biomass CLC


process, the intention of biomass CLG is to produce synthesis gas as
much as possible. Thus, it is of great importance to keep the OC-fuel
ratio at a low value to prevent the complete oxidation of fuel to CO2 and
H2O. In the present study, changing the amount of oxygen carrier
particles may affect the fluidization inside the fuel reactor. Therefore,
the influences of OC-fuel ratio on CLG were investigated by altering the
biomass feeding rate from 0.09 kg/h to 0.18 kg/h, while the inventory
of bed materials was kept as a constant (150 g), with the temperature of
840 °C and argon as the fluidizing agent.
The compositions and total yield of syngas, along with the carbon
conversion efficiency (ηc), gasification efficiency (η) and LHV as the
function of biomass feeding rate are shown in Fig. 9. The CO and H2
concentrations increase smoothly while the CO2 concentration declines
rapidly with the escalation of biomass feeding rate. This can be Fig. 9. Effects of biomass feeding rate on biomass CLG performance.

8
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

gases is considered to be the optimum steam content in this study. As a


consequence of the gas distribution evolution, the LHV of the gas pro-
duct shows a small decrease and is retained at around 14.2 MJ/Nm3
with the increase of steam content.
The evolution of oxygen carrier conversion X and conversion rate
dX/dt with various biomass feeding rate are displayed in Fig. 12. As can
be seen, the conversion of oxygen carriers is significantly enhanced
owing to the substantial growth of the combustible gases yield with the
addition of steam as gasification agent. The time variations of their
differentiation also suggest the faster consumption rate of natural he-
matite as the reactant in the atmosphere with higher steam ratio.

4. Conclusion

A transient CFD model for simulation of biomass chemical looping


Fig. 10. Evolution of OC conversion at different biomass feeding rate. gasification with continuous feedstock in a bubbling fluidized bed fuel
reactor has been developed. By integrating the Eulerian multiphase
(R6–R9) between natural hematite and the larger amount of combus- model with the chemical kinetic models including pyrolysis of biomass,
tible gases generated from biomass pyrolysis and char gasification. The gasification of char, WGS reaction and OC reduction reactions, this
time variations of their differentiation also indicate that the OC con- model is able to provide a relatively good prediction of the time-varying
version rate was firstly accelerated with the increase of biomass feeding outlet concentrations of five gas components in comparison with the
rate but then descended faster due to the rapid consumption of the measured values from the experimental study [21]. The results de-
reactant natural hematite. monstrate the continuous feedstock of pine sawdust and the mixing and
segregation behaviors between fuel and OC particle phases, which have
strong impacts on the gas composition distribution as well as the evo-
3.4. Effects of steam content lution of the solid particles in the CLG system.
The effects of various operation temperature, biomass feeding rate
Steam is considered as an effective gasification agent in the CLG and steam content on the biomass CLG performance are analyzed. It is
process to promote the thermochemical conversion of solid fuel into found that increasing the temperature facilitates the endothermic re-
valuable gaseous products, especially for the generation of H2-rich actions, leading to an increase of CO and H2 concentrations and the
syngas. Moreover, according to the previous study [33], the presence of decrease of CH4, C2H4 and CO2 concentrations. The gas yield and ga-
steam can efficiently prevent Fe3O4 from being further reduced to FeO/ sification efficiency also increase with the escalating temperature but
Fe, which is beneficial for keeping the reactivity of the oxygen carriers the change in LHV is indistinctive as a result of the varying gas com-
[15]. Additionally, the use of water vapor as gasification agent is more positions. Increasing the biomass feeding rate, on the other hand, gives
in line with the needs of industrial fluidized beds [50]. Hence, the in- a lower OC-fuel ratio, resulting in higher concentrations of the com-
fluence of steam content on the biomass CLG performance was in- bustible gases and lower concentrations of CO2 and H2O. Hence, the
vestigated by varying the proportion of steam in the range of 0–75% in LHV and gasification efficiency smoothly ascends but the carbon con-
the fluidizing gas, whose total flow rate was kept at 1500 L/h, with the version efficiency slightly descends due to the less sufficient lattice
fuel feeding rate of 0.12 kg/h and temperature of 840 °C. oxygen as larger amount of biomass enters the reactor. The addition of
Fig. 11 shows the compositions and total yield of syngas along with steam as the gasification agent significantly improves the biomass CLG
the carbon conversion efficiency (ηc), gasification efficiency (η) and performance in terms of H2 concentration, gas yield, carbon conversion
LHV as the function of steam content. As can be seen, the H2 con- efficiency as well as the gasification efficiency. No obvious enhance-
centration increases from 20.83% in pure argon atmosphere to 29.84% ment is observed with the introduction of excess steam, thus 50% in the
with 10% of steam contained in the fluidizing agents, and then fluidizing gases is considered to be the optimum steam content in order
smoothly climbs to 31.86% as the steam content reaches 75% mainly to avoid the potential energy consumption.
due to the steam gasification reaction (R4). Correspondingly, the con-
centrations of CH4, C2H4, CO display a slight downtrend with the in-
creasing steam ratio. Such results are consistent with the previous work
[33] and other publications [23,50]. With the increase of steam con-
tent, the reaction rate of WGS (R5) increases, leading to an escalation in
the mole fraction of H2 and CO2 in the outlet. However, the increment
of CO2 is smaller than the increments of CO and H2 owing to the char
gasification reactions, hence the relative concentration of CO2 shows a
descent firstly but then increases slightly as more steam is introduced
into the fluidized bed reactor.
It is observed that the gas yield is elevated with the increasing steam
content and maintained at around 1.28 Nm3/kg as the steam content
exceeds 50% in the fluidizing gases. Simultaneously, both the gasifi-
cation efficiency and carbon conversion efficiency present an apparent
ascent and then becomes stable at around 91.0% and 99.5%, respec-
tively. Moreover, several studies have pointed out that excess steam
would increase the energy penalty and promote the conversion of CO
into CO2 [15,23,50], indicating that further increasing the steam ratio
in the fluidizing gases may not lead to an improvement but a dimin-
ishment on the biomass CLG performance, thus 50% in the fluidizing Fig. 11. Effects of steam content on biomass CLG performance.

9
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

Chem Eng J 2016;286:689–700. https://doi.org/10.1016/j.cej.2015.11.008.


[15] Liu G, Liao Y, Wu Y, Ma X, Chen L. Characteristics of microalgae gasification
through chemical looping in the presence of steam. Int J Hydrogen Energy
2017;42:22730–42. https://doi.org/10.1016/j.ijhydene.2017.07.173.
[16] Zeng J, Xiao R, Zhang S, Zhang H, Zeng D, Qiu Y, et al. Identifying iron-based
oxygen carrier reduction during biomass chemical looping gasification on a ther-
mogravimetric fixed-bed reactor. Appl Energy 2018;229:404–12. https://doi.org/
10.1016/j.apenergy.2018.08.025.
[17] Hu J, Li C, Guo Q, Dang J, Zhang Q, Lee DJ, et al. Syngas production by chemical-
looping gasification of wheat straw with Fe-based oxygen carrier. Bioresour Technol
2018;263:273–9. https://doi.org/10.1016/j.biortech.2018.02.064.
[18] Shen T, Ge H, Shen L. Characterization of combined Fe-Cu oxides as oxygen carrier
in chemical looping gasification of biomass. Int J Greenh Gas Control
2018;75:762–9. https://doi.org/10.1016/j.ijggc.2018.05.021.
[19] Yin S, Shen L, Dosta M, Hartge EU, Heinrich S, Lu P, et al. Chemical looping ga-
sification of a biomass pellet with a manganese ore as an oxygen carrier in the
fluidized bed. Energy Fuels 2018;32:11674–82. https://doi.org/10.1021/acs.
energyfuels.8b02849.
[20] Zeng J, Xiao R, Zhang H, Chen X, Zeng D, Ma Z. Syngas production via biomass self-
moisture chemical looping gasification. Biomass Bioenergy 2017;104:1–7. https://
doi.org/10.1016/j.biombioe.2017.03.020.
Fig. 12. Evolution of OC conversion at different steam content. [21] Huang Z, He F, Feng Y, Zhao K, Zheng A, Chang S, et al. Synthesis gas production
through biomass direct chemical looping conversion with natural hematite as an
oxygen carrier. Bioresour Technol 2013;140:138–45. https://doi.org/10.1016/j.
Declaration of Competing Interest biortech.2013.04.055.
[22] Wei G, He F, Huang Z, Zheng A, Zhao K, Li H. Continuous operation of a 10 kWth
The authors declare that they have no known competing financial chemical looping integrated fluidized bed reactor for gasifying biomass using an
iron-based oxygen carrier. Energy Fuels 2015;29:233–41. https://doi.org/10.1021/
interests or personal relationships that could have appeared to influ- ef5021457.
ence the work reported in this paper. [23] Ge H, Guo W, Shen L, Song T, Xiao J. Biomass gasification using chemical looping in
a 25 kWth reactor with natural hematite as oxygen carrier. Chem Eng J
2016;286:174–83. https://doi.org/10.1016/j.cej.2015.10.092.
Acknowledgements [24] Larsson A, Israelsson M, Lind F, Seemann M, Thunman H. Using ilmenite to reduce
the tar yield in a dual fluidized bed gasification system. Energy Fuels
The National Research Foundation Singapore, Sembcorp Industries 2014;28:2632–44. https://doi.org/10.1021/ef500132p.
[25] Wang S, Lu H, Zhao F, Liu G. CFD studies of dual circulating fluidized bed reactors
Ltd., and National University of Singapore under the Sembcorp-NUS for chemical looping combustion processes. Chem Eng J 2014;236:121–30. https://
Corporate Laboratory support the research. doi.org/10.1016/j.cej.2013.09.033.
[26] Parker JM. CFD model for the simulation of chemical looping combustion. Powder
Technol 2014;265:47–53. https://doi.org/10.1016/j.powtec.2014.01.027.
References [27] May J, Alobaid F, Ohlemüller P, Stroh A, Ströhle J, Epple B. Reactive two–fluid
model for chemical–looping combustion – simulation of fuel and air reactors. Int J
[1] Xu D, Zhang Y, Hsieh TL, Guo M, Qin L, Chung C, et al. A novel chemical looping Greenh Gas Control 2018;76:175–92. https://doi.org/10.1016/j.ijggc.2018.06.023.
partial oxidation process for thermochemical conversion of biomass to syngas. Appl [28] Luo H, Lin W, Song W, Li S, Dam-Johansen K, Wu H. Three dimensional full-loop
Energy 2018;222:119–31. https://doi.org/10.1016/j.apenergy.2018.03.130. CFD simulation of hydrodynamics in a pilot-scale dual fluidized bed system for
[2] Yao Z, You S, Ge T, Wang CH. Biomass gasification for syngas and biochar co- biomass gasification. Fuel Process Technol 2019;195:106146https://doi.org/10.
production: energy application and economic evaluation. Appl Energy 1016/j.fuproc.2019.106146.
2018;209:43–55. https://doi.org/10.1016/j.apenergy.2017.10.077. [29] Yang S, Wang H, Wei Y, Hu J, Chew JW. Particle-scale modeling of biomass gasi-
[3] Claude V, Courson C, Köhler M, Lambert SD. Overview and essentials of biomass fication in the three-dimensional bubbling fluidized bed. Energy Convers Manage
gasification technologies and their catalytic cleaning methods. Energy Fuels 2019;196:1–17. https://doi.org/10.1016/j.enconman.2019.05.105.
2016;30:8791–814. https://doi.org/10.1021/acs.energyfuels.6b01642. [30] Yang S, Wang H, Wei Y, Hu J, Chew JW. Numerical investigation of bubble dy-
[4] Yang S, Zhou T, Wei Y, Hu J, Wang H. Influence of size-induced segregation on the namics during biomass gasification in a bubbling fluidized bed. ACS Sustain Chem
biomass gasification in bubbling fluidized bed with continuous lognormal particle Eng 2019;7:12288–303. https://doi.org/10.1021/acssuschemeng.9b01628.
size distribution. Energy Convers Manage 2019;198:111848https://doi.org/10. [31] Ostermeier P, Fischer F, Fendt S, DeYoung S, Spliethoff H. Coarse-grained CFD-DEM
1016/j.enconman.2019.111848. simulation of biomass gasification in a fluidized bed reactor. Fuel
[5] Guizani C, Jeguirim M, Gadiou R, Escudero Sanz FJ, Salvador S. Biomass char ga- 2019;255:115790https://doi.org/10.1016/j.fuel.2019.115790.
sification by H2O, CO2 and their mixture: evolution of chemical, textural and [32] Wang S, Yin W, Li Z, Yang X, Zhang K. Numerical investigation of chemical looping
structural properties of the chars. Energy 2016;112:133–45. https://doi.org/10. gasification process using solid fuels for syngas production. Energy Convers Manage
1016/j.energy.2016.06.065. 2018;173:296–302. https://doi.org/10.1016/j.enconman.2018.07.043.
[6] Heidenreich S, Foscolo PU. New concepts in biomass gasification. Prog Energy [33] Li Z, Xu H, Yang W, Xu M, Zhao F. Numerical investigation and thermodynamic
Combust Sci 2015;46:72–95. https://doi.org/10.1016/j.pecs.2014.06.002. analysis of syngas production through chemical looping gasification using biomass
[7] Bridgwater AV. The technical and economic feasibility of biomass gasification for as fuel. Fuel 2019;246:466–75. https://doi.org/10.1016/j.fuel.2019.03.007.
power generation. Fuel 1995;74:631–53. https://doi.org/10.1016/0016-2361(95) [34] Yang S, Wang H, Wei Y, Hu J, Chew JW. Eulerian-Lagrangian simulation of air-
00001-L. steam biomass gasification in a three-dimensional bubbling fluidized gasifier.
[8] Li XT, Grace JR, Lim CJ, Watkinson AP, Chen HP, Kim JR. Biomass gasification in a Energy 2019;181:1075–93. https://doi.org/10.1016/j.energy.2019.06.003.
circulating fluidized bed. Biomass Bioenergy 2004;26:171–93. https://doi.org/10. [35] Ansys. Ansys fluent 12.0 Theory Guide. 2009. DOI:10.1016/0140-3664(87)
1016/S0961-9534(03)00084-9. 90311-2.
[9] Liu G, Liao Y, Wu Y, Ma X. Application of calcium ferrites as oxygen carriers for [36] Li XK. Multiphase flow and fluidization, continuum and kinetic theory descriptions.
microalgae chemical looping gasification. Energy Convers Manage J Nonnewton Fluid Mech 1994;55:207–8. https://doi.org/10.1016/0377-0257(94)
2018;160:262–72. https://doi.org/10.1016/j.enconman.2018.01.041. 80007-3.
[10] Hu Z, Jiang E, Ma X. Microwave pretreatment on microalgae: effect on thermo- [37] Wen CY, Yu YH. Mechanics of fluidization. Chem Eng Progress, Symp Ser
gravimetric analysis and kinetic characteristics in chemical looping gasification. 1966;62:100–11. https://doi.org/10.1016/S0032-0633(98)00014-2.
Energy Convers Manage 2018;160:375–83. https://doi.org/10.1016/j.enconman. [38] Ergun S. Fluid flow through packed columns. Chem Eng Prog 1952;48:89–94.
2018.01.057. doi:citeulike-article-id:7797897.
[11] Salkuyeh YK, Saville BA, MacLean HL. Techno-economic analysis and life cycle [39] Lun CKK, Savage SB, Jeffrey DJ, Chepurniy N. Kinetic theories for granular flow:
assessment of hydrogen production from different biomass gasification processes. Inelastic particles in Couette flow and slightly inelastic particles in a general
Int J Hydrogen Energy 2018;43:9514–28. https://doi.org/10.1016/j.ijhydene. flowfield. J Fluid Mech 1984;140:223–56. https://doi.org/10.1017/
2018.04.024. S0022112084000586.
[12] Adams TA, Barton PI. Combining coal gasification and natural gas reforming for [40] Syamlal M, O’Brien TJ. The derivation of a drag coefficient formula from velocity-
efficient polygeneration. Fuel Process Technol 2011;92:639–55. https://doi.org/10. voidage correlations. West Virginia: US Dept Energy, Off Foss Energy, Natl Energy
1016/j.fuproc.2010.11.023. Tech- Nol Lab Morgantown; 1987.
[13] Zhao X, Zhou H, Sikarwar VS, Zhao M, Park A-HA, Fennell PS, et al. Biomass-based [41] Gunn DJ. Transfer of heat or mass to particles in fixed and fluidised beds. Int J Heat
chemical looping technologies: the good, the bad and the future. Energy Environ Sci Mass Transf 1978;21:467–76. https://doi.org/10.1016/0017-9310(78)90080-7.
2017;10:1885–910. https://doi.org/10.1039/C6EE03718F. [42] Gerber S, Behrendt F, Oevermann M. An Eulerian modeling approach of wood
[14] Ge H, Guo W, Shen L, Song T, Xiao J. Experimental investigation on biomass ga- gasification in a bubbling fluidized bed reactor using char as bed material. Fuel
sification using chemical looping in a batch reactor and a continuous dual reactor. 2010;89:2903–17. https://doi.org/10.1016/j.fuel.2010.03.034.

10
Z. Li, et al. Energy Conversion and Management 201 (2019) 112143

[43] Boroson ML, Howard JB, Longwell JP, Peters WA. Product yields and kinetics from [47] Mahalatkar K, Kuhlman J, Huckaby ED, O’Brien T. CFD simulation of a chemical-
the vapor phase cracking of wood pyrolysis tars. AIChE J 1989;35:120–8. https:// looping fuel reactor utilizing solid fuel. Chem Eng Sci 2011;66:3617–27. https://
doi.org/10.1002/aic.690350113. doi.org/10.1016/j.ces.2011.04.025.
[44] Everson RC, Neomagus HWJP, Kasaini H, Njapha D. Reaction kinetics of pulverized [48] Armstrong LM, Gu S, Luo KH. Effects of limestone calcination on the gasification
coal-chars derived from inertinite-rich coal discards: gasification with carbon di- processes in a BFB coal gasifie. Chem Eng J 2011;168:848–60. https://doi.org/10.
oxide and steam. Fuel 2006;85:1076–82. https://doi.org/10.1016/j.fuel.2005.10. 1021/ie1023029.
016. [49] Wei G, Wang H, Zhao W, Huang Z, Yi Q, He F, et al. Synthesis gas production from
[45] Bustamante F, Enick RM, Cugini A, Killmeyer R, Howard BH, Rothenberger KS, chemical looping gasification of lignite by using hematite as oxygen carrier. Energy
et al. Kinetics of the homogeneous reverse water-gas shift reaction at high tem- Convers Manage 2019;185:774–82. https://doi.org/10.1016/j.enconman.2019.01.
perature. AIChE J 2004;50:1028–41. 096.
[46] Abad A, Adánez J, García-Labiano F, de Diego LF, Gayán P, Celaya J. Mapping of [50] Huang Z, Xu G, Deng Z, Zhao K, He F, Chen D, et al. Investigation on gasification
the range of operational conditions for Cu-, Fe-, and Ni-based oxygen carriers in performance of sewage sludge using chemical looping gasification with iron ore
chemical-looping combustion. Chem Eng Sci 2007;62:533–49. https://doi.org/10. oxygen carrier. Int J Hydrogen Energy 2017;42:25474–91. https://doi.org/10.
1016/j.ces.2006.09.019. 1016/j.ijhydene.2017.08.133.

11

You might also like