Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Energy Conversion and Management 233 (2021) 113863

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Experiment and numerical analysis of catalytic CO2 methanation in


bubbling fluidized bed reactor
Son Ich Ngo a, Young-Il Lim a, *, Doyeon Lee b, Myung Won Seo b, Sungwon Kim c
a
Center of Sustainable Process Engineering (CoSPE), Department of Chemical Engineering, Hankyong National University, Gyeonggi-do, Anseong-si, Jungang-ro 327,
17579, Republic of Korea
b
Climate Change Research Division, Korea Institute of Energy Research, Daejeon, Yuseong-gu, Gajeong-ro 152, 34129, Republic of Korea
c
Department of Chemical and Biological Engineering, Korea National University of Transportation, 50 Daehak-ro, Chungju, Chungbuk 27469, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Keywords: This study experimentally and numerically investigated the hydrodynamics, reaction kinetics, and heat transfer
Power-to-Gas (PtG) of a bench-scale bubbling fluidized bed (BFB) reactor for CO2 methanation. A three-dimensional gas–solid
CO2 methanation Eulerian computational fluid dynamics (CFD) model coupled with a modified Syamlal–O’Brien drag model and
Bubbling fluidized bed (BFB)
reaction kinetics for Ni-based catalysts was developed. The CFD model was validated against experimental data
Feed dilution
Computational fluid dynamics (CFD)
for pressure, temperature, and gas composition at 1 bar and an inlet flow rate of 2 L/min with an inlet N2 content
Heat transfer coefficient (HTC) of 77.5%. The axial pressure drop, solid volume fraction, temperature, gas composition, and bed-to-wall heat
transfer coefficient (HTC) were compared for four inlet N2 contents: 77.5%, 50%, 25%, and 0%. As the inlet N2
content decreased, the mean bed temperature increased from 340 to 456 ◦ C, the gas volume decreased owing to
the reaction, the fluidizing number (ug/umf) decreased from 4.1 to 3.5, and the solid holdup increased. Conse­
quently, the HTC increased from 327 to 386 W/m2/K. This study identified successfully the effects of hydro­
dynamics and reaction kinetics on HTC in the BFB for CO2 methanation.

exchange [12]. However, the design and scale-up of BFB reactors,


especially for highly exothermic reactions such as CO2 methanation,
1. Introduction
require the knowledge of heat transfer between the fluidized medium
and the wall [12] and the influence of the feed concentration on the
Chemical intermediate storage of electricity is one of the most viable
hydrodynamics, bed temperature, and heat transfer [13].
alternatives to balance the time-variant supply and demand of electrical
For non-reactive flows, the heat transfer of a BFB mainly depends on
energy from renewable sources, such as wind and solar energy [1,2].
the inlet gas velocity, particle size, solid concentration, and fluid prop­
The excess energy from renewable sources can be converted to
erties [14,15]. The heat transfer of a reactive BFB is more complex than
combustible chemicals that are easier to store and transport than elec­
that of a non-reactive BFB because of the influence of the reaction heat
trical energy via power-to-gas technology [1,3]. Among several poten­
on the hydrodynamics and heat transfer [13]. Kopyscinski et al. (2011)
tial gaseous energy carriers, hydrogen (H2) and methane (CH4) have a
examined the bed temperature and producer gas composition according
good storage capacity combined with high charge/discharge periods
to the catalyst weight and feed composition in a BFB reactor for syngas
[4]. CH4 has several advantages over H2 in terms of transportation and
methanation [13]. Basu (2015) presented the important factors influ­
storage [1]. Furthermore, the combination of carbon capture and stor­
encing the bed-to-wall heat transfer in a fluidized-bed boiler, such as
age in existing power plants [5,6] with CO2 methanation can recycle
solid concentration, particle size, inlet gas velocity, bed height, and bed
CO2 as a supplementary fuel [7,8].
temperature [15]. Two kinds of the heat transfer coefficient (HTC) were
Among several common catalytic reactors for CO2 methanation, such
reported for BFBs: bed-to-wall HTC caused by convective gas and solid
as fixed-beds and fluidized-beds [3], bubbling fluidized bed (BFB) re­
behaviors [16,17] and bed-to-tube HTC for immersed heat exchangers
actors have advantages in terms of good solid and gas mixing and heat
with coolant [14,18]. Martin (1984) reported that the range of the bed-
transfer [9], uniform temperature [10], low catalyst requirement [3,11],
to-wall HTC in a non-reactive BFB was 200–800 W/m2/K [17]. Kim and
and flexible temperature control using internal or external heat

* Corresponding author.
E-mail addresses: ngoichson@hknu.ac.kr (S. Ich Ngo), limyi@hknu.ac.kr (Y.-I. Lim), dylee82@kier.re.kr (D. Lee), mwseo82@kier.re.kr (M. Won Seo), kswcfb@ut.
ac.kr (S. Kim).

https://doi.org/10.1016/j.enconman.2021.113863
Received 14 November 2020; Accepted 13 January 2021
Available online 24 February 2021
0196-8904/© 2021 Elsevier Ltd. All rights reserved.
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Nomenclature x Thickness m
yi Mole fraction of species i mol/mol
A Surface area of cell element m2 Yi Mass fraction of species i kg/kg
Ai Interfacial area ( = 6αg αs /dp ), m2 /m3 z Transversal direction coordinate m
Ci Concentration of species i mol/m3
Symbols
CD Drag coefficient−
α Volume fraction −
cp Specific heat J/kg/K
βi Adsorption constants of species i 1/bar
db Bubble diameter m ε Relative error−
dp Particle diameter m
γ Energy collisional dissipation rate of solid phase kg/m/s3
D Diffusivity m2/s
κ Thermal conductivity W/m/K
ess Coefficient of restitution (=0.9) −
λ Bulk viscosity Pa∙s
g0,ss Radial distribution coefficient −
μ Shear viscosity Pa∙s
h Height, specific enthalpy, or heat transfer coefficient m, J/
νij Stoichiometric coefficient of ith species in jth reaction −
kg, or W/m2/K
= ϕ Angle of internal friction radian
I Identity tensor − Φ Representative quantity
k Kinetic energy per unit mass J/kg ρ Density kg/m3
K1 , K2 , K3 Equilibrium constants bar2
Stress–strain tensor Pa
=
τ
Kgs Drag coefficient kg/m3 /s
Θ Granular temperature m2 /s2
N Number of cells −
NT Number of time steps − Superscripts
Nus Nusselt number of solid − A Area-averaged
p Order of the discretization method − in Inlet
pi Partial pressure of species ibar out Outlet
P Pressure Pa V Volume-averaged
Pr Prandtl number (=cp μ/κ) −
q Heat flux W/m2 Subscripts
Qgs Heat exchange between gas and solid phase W/m3 b Bubble or bed
QR Reaction heat W col Collision
r Radial coordinate or refinement ratio m or − fri Friction
R1 , R2 , R3 Reaction rates mol/kg/s g Gas-phase
R Reactor radius m i Species index or grid structure index
Rg Gas constant (=8.314) J/mol/K j Reaction index
kin Kinetic
Res Solid particle Reynolds number (=ug ρg dp /μ) −
mf Minimum fluidization
SR Reaction heat source W/m3
q Phase index (g or s)
t Flow time s
R Reaction
T Temperature ◦ C or k
s Solid-phase
u Velocity m/s
t Sampling time
U Overall heat transfer coefficient W/m2 /K w Wall
V Volume m3

Kim (2013) presented that the bed-to-wall HTC of a pressurized BFB rates [3]. The validity of the TPRM depended on the empirical equations
exhibited a maximum value with a variation of the fluidizing number used for the estimation of the hydrodynamic parameters. The geomet­
(ug/umf), where the maximum HTC was 400–500 W/m2/K according to rical effects on the local hydrodynamics and heat transfer of bubbly
pressure [16]. Vogtenhuber et al. (2019) investigated the effect of hor­ flows were not considered in TEM and TPRM.
izontal tube arrangements on the heat transfer from a non-reactive BFB Although the CFD model generally requires more computational
to immersed tubes with the bed-to-tube HTC range of 90–200 W/m2/K time than TEM and TPRM, the knowledge of the local hydrodynamics of
[14]. The bed-to-tube HTC in a bench-scale BFB for CO2 methanation reactive bubbly flows can elucidate the impact of geometrical and/or
was approximately 115 W/m2/K [18]. However, the bed-to-wall HTC of operational modifications [25]. Phuakpunk et al. (2018) presented a
a reactive BFB for CO2 methanation needs to be studied. two-dimensional (2D) Eulerian CFD model coupled with the kinetic
Several mathematical models have been proposed to formulate a theory of granular flow (KTGF) of a circulating fluidized-bed (CFB) for
fluidized-bed with reactions, such as thermodynamic equilibrium steam methane reforming (SMR) that is a reversed reaction of CO2
models (TEM) for biomass gasification [19,20], one-dimensional (1D) methanation, where the riser diameter and operating conditions were
two-phase reaction models (TPRM) for CO and CO2 methanation re­ optimized [26]. Ngo et al. (2019) obtained the overall HTC in a SMR
actions [3,11,21], and computational fluid dynamics (CFD) models for reactor for H2 production using a 3D CFD with chemical reactions [27].
syngas production [22–24]. The TEM is useful for a prompt evaluation of Liu and Hinrichsen (2014) investigated the hydrodynamics and re­
the effect of the most important parameters on the process performance action kinetics in an isothermal BFB reactor for syngas methanation over
[19]. However, the TEM was modified with nonequilibrium factors a Ni/γ-Al2O3 catalyst using a 2D CFD model coupled with KTGF for the
because practical fluidized beds, such as coal and biomass gasifiers, were solid-phase momentum equation [22]. Sun et al. (2018) conducted a 2D
far from thermodynamic equilibrium [19,20]. The TPRM under KTGF-CFD simulation of a CFB for syngas methanation, assuming an
isothermal conditions accounted for both hydrodynamic parameters (e. adiabatic condition [23]. Li et al. (2019) proposed a 2D Euler­
g., size, velocity, and volume fraction of the bubble phase) and reaction ian–Lagrangian CFD model coupled with a discrete element method

2
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

(DEM) in a BFB for syngas methanation [24]. Fattahi et al. (2019)


analyzed the bed-to-wall HTC obtained from a KTGF-CFD model of a
BFB without reaction [28]. Nevertheless, studies on a CFD model of
reactive bubbly flows for CO2 methanation are scarce. No studies have
addressed the experimental validation of a CFD model including heat
transfer for CO2 methanation. Moreover, the effect of the feed dilution
by recirculation [3,29] or inert gas [21] on heat transfer was not
investigated for BFBs for CO2 methanation, to our knowledge.
In this study, CO2 methanation over a Ni/Al2O3 catalyst (Geldart B
particles) in a bench-scale BFB reactor was investigated experimentally
and numerically. The dilute feed gases CO2, H2, and N2 were injected
into the BFB reactor for producing CH4 at the rate of 5 L/h at 25 ◦ C. The
axial pressure drops, temperatures, and gas concentrations were
experimentally measured and used for the validation of the CFD results.
A KTGF-CFD model was employed to investigate the hydrodynamics,
reaction kinetics, and heat transfer of the BFB for CO2 methanation.
After the CFD model was verified and validated, the effects of inlet gas
compositions with the dilution of 0–77.5 vol% N2 on the hydrody­
namics, temperature, and HTC were examined in the BFB reactor.

2. Experimental setup

A green methane catalyst (GMC-Y3) was used as the catalyst for CO2
methanation in the BFB reactor. The GMC-Y3 catalyst, including 50 wt% Fig. 1. Schematic diagram of the experimental setup of the bench-scale BFB
NiO, was supported by a mixture of Al2O3, TiO2, and SiO2. After calci­ reactor for CO2 methanation (MF: mass flowmeter, TC: thermocouple, DP:
nation, the catalyst was sieved to obtain a controlled particle size differential pressure, PC: personal computer).
ranging from 36 to 161 μm (mean diameter of 105 μm). The catalyst
belongs to the Geldart group B. Table 1 summarizes the properties of the 0.0128 m/s without reaction, as shown in Fig. 2. The minimum fluid­
catalyst used in this study. ization velocity (umf ) was determined to be 0.0076 m/s, which is
Fig. 1 shows the schematic diagram of the bench-scale experimental observed at the peak of ΔP.
setup consisting of a BFB reactor, gas injection system, electric pre­
heater, stepping motor for the temperature measurement along the 3. Modeling of BFB reactor for CO2 methanation
height, and on-line gas chromatograph (Advance Optima Uras-14,
Hartmann & Braun Co.) for analyzing the composition of the product The computational domain of the experimental BFB reactor was
gas. restricted to the fluid area of the BFB. The gas–solid Eulerian CFD model
The BFB reactor was made of stainless steel (SUS 310) with an inner in the three-dimensional (3D) computational domain included the
diameter of 0.05 m and a height of 0.68 m. The reactor was initially conversation of mass, momentum, and energy. The inlet boundary
filled with catalyst particles up to a height of 0.25 m and maintained a condition of the CFD model was set to the same value as the experi­
reaction temperature by the electric preheater. With a linear step-motor, mental condition.
the thermocouple (TC) probe was immersed axially from the top to the
catalyst bed. The differential pressure (DP) was measured at heights of 3.1. Computational domain and meshing
0.02 m, 0.06, and 0.49 m from the bottom. Dilute feed gas with a total
flow rate of 2 L/min at 25 ◦ C was injected into the reactor under ambient The computational domain of the BFB reactor is shown in Fig. 3a,
pressure. The feed compositions were 4.5 vol%, 18 vol%, and 77.5 vol%
for CO2, H2, and N2, respectively. N2 was used as an inert gas to mitigate
the exothermic reaction. The electric power was not supplied to the
reactor after feeding the reactant. The product gas exiting the reactor
was condensed to remove water and analyzed using a continuous non-
dispersed infrared (NDIR) gas analyzer (Advance Optima, Hartmann &
Braun Co.).
The total pressure drop (ΔP) of the catalyst bed was measured by
increasing the inlet superficial gas velocity (uinlet) of N2 from 0.0039 to

Table 1
Properties of the catalyst (GMC-Y3) used in this study.
Physical properties Particle true density (kg/m3) 3,905
Apparent density (kg/m3) 2,100
Particle mean diameter (µm) 105
Pore volume (m3/g) 0.01
Sphericity ≈1
Inductively coupled plasma analysis Ni (%) 46.3
Al (%) 13.0
Ti (%) 8.4
Si (%) 2.9
Ca (%) 2.9
Others (%) 26.5
Total (%) 100
Fig. 2. Minimum fluidization velocity (umf ) measured in the bench-scale BFB.

3
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Fig. 3. Computational domain and meshing of the BFB reactor for CO2 methanation.

where the wall thickness and insulation layer were replaced by the wall and Reynolds numbers (Res up to 105) in Eq. (T20); and (9) the colli­
boundary condition. The inlet gas enters at the bottom, and the outlet sional dissipation of kinetic energy (γs) due to inelastic particle collision
gas exits from the top of the BFB through a narrow tube with a diameter in Eq. (T24).
of 8 mm. The initial bed height (Hinit), voidage, and catalyst mass were The drag coefficient (CD) in Eq. (T7) plays an important role in the
0.25 m, 0.435, and 0.58 kg, respectively. A polyhedral mesh structure momentum transfer between the gas and solid phases [25]. The Syam­
derived from unstructured tetrahedral meshes is shown in Fig. 3b for lal–O’Brien drag model [36] was based on the measurements of the
coarse, medium, and fine meshes with 29,100, 108,620, and 180,150 terminal velocity of particles (ur,s) as a function of two coefficients A and
cells, respectively. The three mesh structures were used for the grid B in Eq. (T9) [39]. Two parameters (c1 and d1) appearing in coefficient B
convergence test in the verification step of the CFD model. were modified in this study to match the minimum fluidization condi­
tion of the BFB for CO2 methanation, yielding umf (=0.0076 m/s, see
3.2. CFD model Fig. 2) and αg,mf (=0.414).
The CO2 methanation kinetics proposed by Xu and Froment (1989)
The Eulerian–Eulerian CFD (EE-CFD) model is favorable for a system [40] were employed:
/
with a dense particle phase such as BFB [30,31], as Lagrangian CFD CO + 3H2 ⇌CH4 + H2 O, ΔHR,1 = − 206.1 kJ mol − CO (1)
models, such as DEM, require a large number of particles, which in­
/
creases the computational time [25]. Table 2 presents an unsteady-state CO2 + H2 ⇌CO + H2 O, ΔHR,2 = +41.2 kJ mol − CO2 (2)
two-phase EE-CFD model with the continuity, momentum, energy,
species transport, and granular temperature equations. /
CO2 + 4H2 →2H2 O + CH4 , ΔHR,3 = − 164.9 kJ mol − CO2 (3)
It was assumed that the gas phase was incompressible and the solid
phase was continuous. The gas density (ρg) was calculated using the The chemical reaction rate (Ri) of species i is expressed as follows:
Soave–Redlich–Kwong equation of state. As the inertial energy loss in ∑3
dyg,i
the dense gas–solid flows is mainly caused by local changes in the flow Ri = ρ g =η νi,j rj , i = CO2 , H2 , CH4 , H2 O, and CO (4)
dt
direction rather than the gas phase turbulence [32], a laminar gas flow j=1

has been assumed in the EE-CFD model for BFBs [33,34]. The solid
particle motion was captured by the KTGF in terms of the granular where νi,j is the stoichiometric coefficient for species i and reaction j. rj is
temperature (Θs ) in Eq. (T10) [35]. the reaction rate [27,40,41] and η is the effectiveness factor accounting
The constitutive relations required to close the governing equations for the diffusional limit of large catalyst particles [3,27]. In this study, η
was fixed to 0.95 [3] because fine particles (dp = 105 μm, see Table 1)
include (1) the stress tensors ( τ g and τ s ) of the gas and solid phases in
= =
were used. The reaction kinetics model (Ri) was solved using a two-step
Eqs. (T3) and (T5), respectively; (2) the momentum exchange coefficient
heterogeneous stiff chemistry solver. In the first step, the species
(Kgs) between the gas and solid phases in Eq. (T6); (3) the solid pressure
transport equation in Eq. (T22) for each species was solved spatially
(Ps) in Eq. (T10) representing the normal solid force due to the parti­
with the reaction rate (rj) set to zero at a given timestep. In the second
cle–particle interaction [31]; (4) the solid bulk viscosity (λs) in Eq. (T11)
step, rj was integrated in every computational cell using a stiff ordinary
accounting for the resistance of the solid phase according to compres­
differential equation solver. The two steps were repeated until the
sion and expansion; (5) the radial distribution coefficient (g0,ss ) in Eq.
temperature and gas composition converged to a tolerable error be­
(T12) expressing the probability of particle collisions when the granular
tween the two steps with a relative tolerance of 1 × 10− 5. The rate (rj)
phase becomes dense; (6) the solid shear viscosity (μs) in Eq. (T13)
was provided with a user-defined function.
representing the tangential force due to the particle interactions
including the collision (μs,col), kinetic (μs,kin), and friction (μs,fri) terms;
(7) the chemical reaction rate (Ri) and heat of reaction (ΔHR,i) in Eq. 3.3. Boundary conditions and model parameters
(T18); (8) the convective gas–solid heat transfer rate (Qgs) with Gunn’s
model (Nus) [38] validated within a wide range of αg values (0.35.1.0) The CFD simulation was performed for four cases with the same inlet

4
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Table 2
Eulerian–Eulerian computational fluid dynamics (EE-CFD) model.
- Continuity equation
( )
∂ ρq αq ( )

u q = 0 where q = g or s
+∇ ∙ ρq αq → (T1)
∂t
-
Momentum equation of the gas phase
( ) ( )
u g) →
∂(ρg αg → → → =
u g→
+ ∇ ∙ ρg αg → ug = − αg ∇ P + ∇ ∙ τ g + ρg αg → g + Kgs (→ug − →u s) (T2)
∂t ( )
→→ →→T 2→ → = (T3)
where the stress tensor of the gas phase is τ g = αg μg ∇ u g +∇ u g − ∇ ∙ u g I
=
3
-
Momentum equation of the solid phase
( )
u s) →
∂(ρs αs → → → → (= )
u s→
+ ∇ ∙ ρs αs → u s = − αs ∇ P − ∇ Ps + ∇ ∙ τ s + ρs αs →g − Kgs (→
ug − → u s) (T4)
∂t ( )
→→ →→T 2→ → = → =
where the stress tensor of the solid phase: τ s = αs μs ∇ u s +∇ u s − ∇ ∙ u s I + αs λs ∇ ∙→
u s I,
=
3 (T5)
( )⃒ ⃒
3 αs αg ρg Res ⃒⃒→ → ⃒⃒
the momentum exchange coefficient: Kgs = CD u g − u s ⃒, (T6)
4 u2r,s dp ur,s ⃒
( )2
4.8
the drag coefficient [36]: CD = 0.63 + √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ,
Res /ur,s (T7)
[ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ]
the terminal velocity: ur,s = 0.5 A − 0.06Res + (0.06Res )2 + 0.12Res (2B − A) + A2 , (T8)

⎨ c1 α1.28 αg ≤ 0.85
g
the coefficients of ur,s: A = α4.14
g ,B= with c1 = 1.51 and d1 = − 1.26,
⎩ αdg 1 αg > 0.85 (T9)

1
the solid-phase pressure: Ps = 2ρs (1 +ess )α2s g0,ss Θs with the granular temperature Θs = us us , (T10)
3
1
4 (Θ )
s 2
the solid-phase bulk viscosity: λs = αs ρs dp g0,ss (1 +ess ) , (T11)
3 π
[ ( )1 ]− 1
αs 3
the radial distribution coefficient [35]: g0,ss = 1 − with αs,max = 0.61, (T12)
αs,max

the solid shear viscosity: μs = μs,col + μs,kin + μs,fri , (T13)

4 (Θ )1
s 2
the collision viscosity of solids [36]: μs,col = αs ρs dp g0,ss (1 +ess ) αs , (T14)
5 π
√̅̅̅̅̅̅̅̅ [ ]2
10ρs dp Θs π 4
the kinetic viscosity of solids [35]: μs,kin = 1 + g0,ss αs (1 + ess ) αs , and (T15)
96αs (1 + ess )g0,ss 5

Ps sinϕ
the frictional viscosity of solids [37]: μs,fri = √̅̅̅̅̅̅̅ with ϕ = 30 .

2 I2D (T16)
-
Energy equation of the gas phase
∂( ) → ( )

(

)
ρ α h + ∇ ∙ ρg α g → u g + SR + Qgs
u g hg = ∇ ∙ α g κ g ∇ T g + τ g ∙ →
=
∂t g g g (T17)

where the heat source from the reaction: SR = 3i=1 Ri ∙ΔHR,i with the reaction rate Ri and the reaction heat ΔHR,i . (T18)
-
Energy equation of the solid phase
( ) ( )
∂ → dPs → →
(ρ αs hs ) + ∇ ∙ ρs αs →
u s hs u s − Qgs
+ ∇ ∙ αs κs ∇ Ts + τ s ∙→
=
∂t s
= αs
dt (T19)

κg Nus ( )
where the heat transfer rate from the gas to solid phase: Qgs = Ai Ts − Tg , and (T20)
dp
( )( ) ( )
1/3 1/3
Gunn’s model [38]:Nus = 7 − 10αg +5α2g 1 +0.7Re0.2
s Pr + 1.33 − 2.4αg +1.2α2g Re0.7
s Pr (T21)
- Species transport equation of the gas phase

( )
∂(ρg αg Yg,i ) → → ( → )
u g Yg,i
+ ∇ ∙ ρg αg → = Dg,i ∇ ∙ ρg αg ∇ Yg,i + Ri (T22)
∂t
-
Granular temperature equation [35]
( ( ))
3 ∂ρs αs Θs → ( = = ) →
u s Θs
+∇ ∙ ρs αs → = − Ps I + τ s : ∇ →
v s − γs − 3Kgs Θs (T23)
2 ∂t ( )
12 1 − e2ss g0,ss 2 3/2
where the energy collisional dissipation rate of the solid phase: γs = √̅̅̅ ρs αs Θs (T24)
dp π

5
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Table 3
Inlet, wall, and outlet boundary conditions for four case studies.
Name Boundary condition Value Remarks

Inlet Inlet N2 content (vol.%) 77.5 (exp.), 50, 25, and 0 H2/CO2 = 4
Velocity (m/s) 0.035 4.6umf
Temperature (◦ C) 340 Exp. condition
Outlet Pressure (bar) 1
Wall Outside wall temperature (◦ C) 150
Shear condition No-slip

flow rate (=2 L/min) and H2/CO2 ratio (=4) by changing the inlet N2 species equations were discretized using the 1st-order upwind method.
content from 77.5 to 0 vol%. The case with 77.5% N2 was the experi­ The 1st-order implicit method was used for the time integration in the
mental condition. The boundary conditions of the inlet, outlet, and wall unsteady-state CFD model. The time step of the transient simulation
are summarized in Table 3. The inlet boundary condition was set to the ranged from 0.0001 to 0.001 s. The maximum number of iterations was
velocity of 0.035 m/s (=4.6umf ) and the temperature of 340 ◦ C, whereas 30 at each timestep, and the convergence tolerance was 1 × 10− 3.
the outlet boundary condition was set to the pressure of 1 bar. No-slip The time-, area-, and volume-averaged quantities are required to
boundary condition was assumed for the wall. The heat flux (qw) lost analyze the temporal- and spatial-dependent physical quantities from
through the reactor wall was estimated using the heat conductions for the CFD results [27,45]. The time-averaged quantity (〈Φ〉) of a variable
the steel and insulator. (Φ) was calculated in the transient simulation as follows:
∑N
1 k=1 Φk Δtk
qw = Δxsteel (T − Two ) (5) 〈Φ〉 = ∑ (6)
κsteel
+ Δx insul
κinsul
N
k=1 Δtk

where Δxsteel (=3 mm) and Δxinsul (=50 mm) are the thicknesses of the where Φk is the value at time step k, and Δt is the timestep size. The time-
steel and insulator, respectively. κsteel (=16 W/m/K) and κinsul (=0.0025 averaged quantities were calculated over a flow time of 40 s, after the
W/m/K) [42] are the thermal conductivities of the steel and insulator, system reached a stable state (t = 200 − 300 s). The area-averaged
respectively. Thus, the heat transfer coefficient (hw) through the wall − A
quantity (Φ ) of the variable Φ for n cells within a cross-sectional area
was 0.5 W/m2/K. The temperature (Two) of the outside wall was set to
(A) was calculated as
150 ◦ C.
∑n
The temperature-dependent physical properties of gas species, such − A Φ k Ak
Φ = k=1 (7)
as density (ρg ), viscosity (μg ), heat capacity (cp,g), conductivity (κg), and A
diffusivity (Dg), have been reported elsewhere [27]. The mixing rule − A
[43] was used to calculate the gas mixture properties. The isotropic The time- and area-averaged quantities are denoted by 〈Φ 〉. The
− V
thermal conductivity of the catalyst (κs) was set to 0.34 W/m/K, which volume-averaged quantity (Φ ) of the variable Φ for m cells within a
was interpolated for the 46.3% Ni catalyst (see Table 1) [44]. volume (V) was calculated as
A CFD commercial code, ANSYS Fluent 19.1 (ANSYS Inc., USA), was ∑m
used to solve the CFD model with a 24-core workstation (Intel Xeon CPU − V
Φ = k=1
Φ k Vk
(8)
E5-2670 at 2.3 GHz and 128 GB RAM). The finite volume method was V
implemented to calculate the 3D partial differential equations, where − V
The time- and volume-averaged quantity is denoted by 〈Φ 〉.
the volume integral of the divergence terms was converted to the surface
integral over the computational cell. The phase-coupled semi-implicit
method for pressure-linked equations (SIMPLE) was used to link the 4. CFD model verification and validation
velocity and pressure of the gas–solid multiphase flow. The quadratic
upwind interpolation for convection kinetics (QUICK) scheme was The CFD model needs to be verified and validated to guarantee the
employed as a 3rd-order upwind method for the first-order spatial dis­ reliability. In the verification step, the grid convergence index (GCI)
cretization of the continuity and momentum equations. The pressure method [46] was adopted to ensure the discretization accuracy related
→ to the mesh size for confirming that the model is accurately solved. In
gradient (∇ P) was discretized using the pressure staggering option
the validation step, the CFD results, such as pressure (P), temperature
(PRESTO) scheme. The first-order spatial gradients of the energy and
(T), and species mole fractions (yi), were compared against experimental

Table 4
Grid convergence calculation using the GCI method for the BFB reactor for CO2 methanation.
Φ i Ni Φi ri,i+1 εi,i+1 GCIi,i+1 (%) ζ

− A 0 ∞ 2,388 1.0004
ΔP (Pa)
1 180,150 2,390 1.184 0.00043 0.09
2 108,616 2,391 1.551 0.00284 0.14
3 29,103 2,398
ug (m/s)
− V 0 ∞ 0.0873 0.9849
1 180,150 0.0852 1.184 − 0.01515 3.16
2 108,616 0.0839 1.551 − 0.09826 5.13
3 29,103 0.0756
− V 0 ∞ 348.9 1.0003
Ts (◦ C)
1 180,150 349.0 1.184 0.00032 0.05
2 108,616 349.1 1.551 0.00244 0.09
3 29,103 350.0

6
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

− A − V − V
Fig. 4. Asymptotic solutions of pressure drop (ΔP ), gas velocity (ug ), and solid temperature (T s ) in the grid convergence test.

Fig. 5. Validation of the CFD results with the experimental data for time-averaged pressure (〈P〉), gas temperature (〈Tg 〉), and species mole fractions (〈yi 〉).

data in the BFB reactor for CO2 methanation to confirm that the model uncertainty in the error estimation. The ratio of the two GCIs (ζ) used as
accurately represents the real world. a criterion of the grid convergence test is
ζ = rp12 GCI12 /GCI23 (13)
4.1. Grid convergence test
When ζ is close to unity, the CFD results of Φ approach the asymp­
totic solution [46], and further grid refinement is not necessary owing to
The GCI method [46] is based on the Richardson extrapolation
both an insignificant improvement of the accuracy and a significant
method involving the comparison of discrete solutions at two different
increase in the calculation time. The exact solution at zero-grid spacing
grids. Three levels of the grid structure, i.e., fine (i = 1), medium (i =
(Φ0 ) from the Richardson extrapolation method is expressed as [46]
2), and coarse (i = 3) meshes (see Fig. 3b), were used to determine the
independency of the key quantities (Φi ), such as pressure drop (ΔP), gas
p
Φ0 = Φ1 + (Φ1 − Φ2 )/(r12 − 1) (14)
velocity (ug), and solid temperature (Ts). The mean grid sizes (Δh =
Table 4 summarizes the grid convergence calculations for the BFB
(VBFB /Ntot )1/3 ) for the BFB reactor volume VBFB and the total number Ntot − A − V
of cells were 1.81 mm, 2.15 mm, and 3.33 mm for the fine, medium, and reactor for CO2 methanation. The ζ values of ΔP and T s are close to
− V
coarse meshes, respectively. The relative error (εi) of Φi between two unity, whereas that of ug is a little far from unity. Roache (1994) re­
levels of grids and the grid refinement ratio (ri) are defined as ported that a GCI of less than 5% is acceptable with a 95% confidence for
εi,i+1 = (Φi+1 − Φi )/Φi (9) the grid convergence test [46].
− A − V − V
Fig. 4 displays the asymptotic solutions of ΔP , us , and Ts in the grid
ri,i+1 = (Ni /Ni+1 )1/3
(10) convergence study, where the solid-lines represent the solutions (Φi, i =
1, 2, and 3) on the three mesh structures and the dashed-lines are the
The order of the discretization method (p) is obtained by solving the
asymptotic solutions (Φ0) at Δh = 0 (zero-grid spacing) obtained from
following non-linear equation [46]:
Eq. (14). The relative errors between Φ0 at Δh = 0 and Φ2 on the medium
p
(Φ3 − Φ2 )/(r23 p
− 1) = r12 p
[(Φ2 − Φ1 )/(r12 − 1) ] (11) − A − V − V
mesh for ΔP , us , and Ts are 0.1%, 3.9%, and 0.1%, respectively. The
The GCI between two adjacent grid structures is calculated as medium mesh (Δh = 2.1 mm) with ζ close to unity represents the
( ⃒ ⃒ )/( p ) asymptotic solutions well. Considering both the calculation time and
GCIi,i+1 (%) = 100 Fs ⃒εi,i+1 ⃒ ri,i+1 − 1 (12)
numerical accuracy, the medium mesh (108,616 cells) was selected for
further investigations.
where the safety factor (Fs = 1.25) was used [46] to consider the

7
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Fig. 6. Time-averaged pressure (〈P〉) for the four inlet N2 contents.

Fig. 7. Time-averaged vertical solid velocity (〈uy,s 〉) for the four inlet N2 contents.

4.2. CFD model validation [13]. As shown in Fig. 5b, the 〈Tg 〉 obtained from the CFD simulation
follows the experimental data along h, which indicates that the heat
The CFD results are compared with the experimental data for the transfer model with the wall boundary condition represents the reaction
time-averaged pressure (〈P〉), gas temperature (〈Tg 〉), and species mole heat generation and wall heat loss.
fractions (〈yi 〉) along the BFB reactor height (h), as shown in Fig. 5. Four In Fig. 5c, the mean values of the predicted mole fractions of the
pressures measured at the gas inlet, DP1, DP2, and DP3 (see Fig. 1a) are species are consistent with the experimental data for the inlet N2 content
shown in Fig. 5a. The pressure profile varies linearly from the distributor of 77.5%. The CO mole fraction was almost zero in the dried product gas
to the bed surface and is constant in the freeboard. The temperature and and the rest of the gas composition shown in Fig. 5c was N2 varying
species mole fractions were measured at 26 points from the distributor between 77.5% and 91% from the bottom to the top of the reactor. The
to the freeboard in the experiment by using a movable device. The gas CO2 conversions obtained from the experiment and CFD simulation were
temperature increases steeply from the inlet (340 ◦ C) to approximately 82% and 87%, respectively. The variation values of 〈yi 〉 (shaded region)
350 ◦ C at h = 0.015 m, and then stabilizes at 349 ◦ C until the bed height are relatively wide because of the bubble dynamics inside the BFB. As
at z = 0.25 m. The BFB shows excellent uniformity of the axial tem­ the temperature profile matches well with the experimental measure­
perature. In the freeboard region, the gas temperature decreases grad­ ment, a small deviation between the CFD results and the experimental
ually because of heat loss through the reactor wall, which is similar to data is observed, except for H2. The consumption of H2 obtained from
the trend for syngas methanation reported by Kopyscinski et al. (2011) the CFD model is greater than that in the experiment. This may result

8
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

Fig. 8. Bubble shape at a flow time of 300 s and time-averaged solid volume fraction (〈αs 〉) for the four inlet N2 contents.

Table 5
Time- and volume-averaged physical properties of gas for the four inlet N2 contents.
Inlet N2 content Temperature Voidage Density Viscosity Specific heat capacity Thermal conductivity
(%) (Tg, ◦ C) (αg) (ρg, kg/m3) (μg, kg/m/s) (cp,g, J/kg/K) (κg, W/m/K)

77.5 348 0.566 1.002 2.97 × 10-5 1239 0.055


50 405 0.574 0.821 3.05 × 10-5 1513 0.071
25 427 0.562 0.680 2.98 × 10-5 1924 0.091
0 456 0.553 0.514 2.85 × 10-5 2682 0.121

Fig. 9. Time-averaged gas temperature (〈Tg 〉) and reaction heat (〈SR 〉) for the four inlet N2 contents.

from the reaction kinetics proposed by Xu and Froment [40], which was 5. Results and discussion
based on a catalyst different from that used in this study. Moreover, as
the polydispersed particles used in the experiment were assumed to be After model verification and validation, the effects of inlet gas
monodispersed in the CFD model, a deviation of yi may occur. compositions on the hydrodynamics, reaction kinetics, and heat transfer
were investigated. In particular, the variation in the HTC inside the BFB
reactor was examined according to the inlet gas composition.

9
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

5.1. Hydrodynamics according to inlet N2 content The total volume of bubbles is directly related to the bed expansion
colored by the magnitude of 〈αs 〉 as shown in Fig. 8b. The bed expansion
The flue gas recirculation was used for the feed dilution in fixed- for 50% inlet N2 content is slightly higher than those for the other inlet
beds, decreasing the hot spot temperature [3,29]. Jia et al. (2020) N2 contents because of the higher volume of bubbles within the catalyst
[21] investigated the effect of operating conditions in a BFB at an inlet bed. The opposite effect of the thermal explosion (temperature increase)
N2 content of 20%. In this subsection, the effects of the feed dilution by and volume contraction (CO2 methanation) on the bed expansion leads
the inlet N2 content on hydrodynamics, heat transfer, and CO2 conver­ to the maximum bed expansion at 50% inlet N2 content (see also
sion are investigated using the CFD model. Table 5).
The contours on the slice normal to the transversal direction (z = 0) Fig. 8c shows the radial profiles of 〈αs 〉 at h = 0.15 m. The 〈αs 〉 at the
and the axial profiles of time-averaged pressure (〈P〉) at the center line center is lower than that at the wall because the solid particles fall near
along the BFB height (h) are compared in Fig. 6 for the four inlet N2 the wall exhibiting core-annular flow. The radial 〈αs 〉 for 0% inlet N2
contents (77.5%, 50%, 25%, and 0%). As the inlet N2 content decreased content was the highest among the four inlet N2 contents, and the bed
from 77.5% to 0%, there was no perceptible change in 〈P〉 decreasing expansion was the lowest. The radial 〈αs 〉 for 77.5% inlet N2 content was
linearly from the bottom to the top. However, 〈P〉 at 0% inlet N2 is the similar to that for 50% inlet N2 content, showing a similar bed
lowest because the contraction of the gas volume by CO2 methanation expansion.
(see Eq. (3)) dominates the thermal expansion due to the exothermal Fig. 9 shows the contours and axial profiles of the time-averaged gas
heat of reaction (see Section 5.2). temperature (〈Tg 〉), and the time-averaged total heat of the three re­
Fig. 7 shows the contours and axial profiles of the time-averaged actions (〈SR 〉) for the four inlet N2 contents. As the BFB exhibits excellent
vertical solid velocity (〈uy,s 〉) for the four inlet N2 contents. The posi­ mixing and heat transfer [11], the axial and radial 〈Tg 〉 within the
tive value represents the upward direction of the particles, whereas the catalyst bed is almost isothermal except for the bottom of the reactor
negative value indicates the downward direction. For core-annular flow where a strong reaction occurs. The temperature increases steeply from
in the fluidized-beds, the solid particles rise at the center and drop near 340 ◦ C to a certain temperature because of the large amount of heat of
the wall. 〈uy,s 〉 is prone to increase from the bottom to the top of the reaction at the bottom. Fig. 9b shows the axial profiles of 〈Tg 〉 at the
catalyst bed, in line with the increase in the bubble rising velocity. As the center line. The mean bed temperatures were 348 ◦ C, 405 ◦ C, 427 ◦ C,
inlet N2 content decreased from 77.5% to 0%, the 〈uy,s 〉 decreased at and 456 ◦ C for the inlet N2 contents of 77.5%, 50%, 25%, and 0%,
most heights of the catalyst bed. This is attributed to the same reason as respectively. It is evident that the gas volume (V2) for 0% inlet N2
that for P in Fig. 6. The influence of the volume contraction by the re­ content is lower than that (V1) for 77.5% inlet N2 content considering
action on the gas velocity is higher than that of the thermal expansion the following relation derived from the ideal gas law and the CO2 con­
due to the temperature increase. The reduced gas velocity decreased version (X):
〈uy,s 〉.
V2 (60X2 + 100(1 − X2 ))T2
Fig. 8 shows snapshots of the bubble shape at a flow time of 300 s, = (15)
V1 (77.5 + 13.5X1 + 22.5(1 − X1 ))T1
and the contours and radial profiles of the time-averaged solid volume
fraction (〈αs 〉) for the four inlet N2 contents. In Fig. 8a, the bubbles are
where T1 and T2 in Kelvin are the bed temperatures at 77.5% and 0%
detected for computational cells having an αs smaller than 0.35. The
inlet N2 contents, respectively. X1 and X2 are 0.874 and 0.801, respec­
bubbles are colored according to the magnitude of the vertical solid
tively, according to the reaction kinetics. V2 is 88% of V1 in Eq. (15).
velocity (uy,s). The larger the bubbles, the higher is the uy,s. The snap­
However, the bed expansion at 0% inlet N2 content was 97% that at
shots at 300 s represent a general feature of the bubble shapes, compared
77.5% inlet N2 content (see Fig. 8b), which is not proportional to the gas
with the six snapshots from 270 to 295 s with an interval of 5 s in Fig. S1
volume reduction. As shown in Fig. 9c, a large amount of 〈SR 〉 is released
of the Supplementary Material. A considerable number of bubbles
at the bottom. The 〈SR 〉 for 0% inlet N2 content is the highest because all
appear for 50% inlet N2 content, whereas the number of bubbles reduces
the components participate in CO2 methanation.
for 0% inlet N2 content.

5.2. Reaction kinetics according to inlet N2 content

Fig. 10 shows the contours and axial profiles of the time-averaged


CH4 mole fraction (〈yCH4 〉) for the four inlet N2 contents. The 〈yCH4 〉
for 0% inlet N2 content was the highest, as expected. 〈yCH4 〉 is relatively
uniform over the reactor, as shown in Fig. 9a. The 〈yCH4 〉 at the center
line increases sharply at the bottom, as shown in Fig. 9b. With the
decrease in inlet N2 content, the CO2 conversion decreased (87.4 vol%,
82.4 vol%, 81.9 vol%, and 80.1 vol%, respectively), because the CO2
methanation is thermodynamically favored by a low-temperature con­
dition [3,40] (see Fig. 9b).

5.3. Heat transfer according to inlet N2 content

− A
Fig. 11 shows the time- and area-averaged heat flux (〈q 〉), temper­
− A − A − A
ature difference (〈Tb − T wi 〉), and bed-to-wall convective HTC (〈hc 〉)
along with the reactor height (h). The time- and area-averaged quanti­
ties were obtained for sleeve-shape segments of the first cell from the
wall [28]. The HTC (hc ) is expressed as follows:
q
hc = (16)
Tb − Twi
Fig. 10. Time-averaged CH4 mole fraction (〈yCH4 〉) for the four inlet
N2 contents. where q (W/m2) is the heat flux from the fluidized-bed to wall, Tb is the

10
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

− A − A − A − A
Fig. 11. Time- and area-averaged heat flux (〈q 〉), temperature difference (〈T b − T wi 〉), and bed-to-wall convective HTC (〈hc 〉) for the four inlet N2 contents.

than N2. Therefore, cp,g and κg at the 0% inlet N2 content are two times
higher than those at the 77.5% inlet N2 content. The high κg enhances
heat transfer in the BFB [47].
Fig. 12 compares the mean bed-to-wall HTC (hc) obtained from
Fig. 11c with an empirical equation [47] and experimental data [16]
with respect to the fluidizing number (ug/umf). The empirical equation
[47] for the Nusselt number (Nu) was proposed for a range of 74 ≤ dp ≤
4,000 μm, 26 ≤ ρs ≤ 11,800 kg/m3, 0.03 ≤ Pg ≤ 2 MPa, 290 ≤ Tb ≤
1,050 K, and ug up to 2.5 m/s. The empirical bed-to-wall hc was calcu­
lated from the Nu with the physical properties reported in Table 5. The
experimental data [16] was referred to as the hc in an air-sand BFB at 1
≤ ug/umf ≤ 4 and 0.1 ≤ P ≤ 0.74 MPa. As discussed previously, the mean
hc obtained from the CFD simulation is higher at the lower fluidizing
number because of the high solid holdup and thermal conductivity of
gas, which coincides with those from the empirical equation and
experimental data. The mean hc for h ≤ 0.2 m are 322, 338, 358, and
386 W/m2/K for the inlet N2 contents of 77.5%, 50%, 25%, and 0%,
respectively. The predicted hc is lower than that of the Molerus’s
empirical equation [47], whereas it is comparable to the experimental
data by Kim and Kim (2013) [16]. As the hc increases with increasing
Fig. 12. Bed-to-wall HTCs (hc) obtained from CFD simulation with reaction, pressure due to the increase of gas density [16], the three experimental
empirical equation, and experiment without reaction. points above 400 W/m2/K, which were measured at 0.21 – 0.74 MPa,
should be higher than the hc predicted at 0.1 MPa. This CFD study
bulk temperature at 3 mm from the wall in the normal direction, and Twi identified successfully the effects of hydrodynamics and reaction ki­
is the inner wall temperature, which were obtained from the CFD netics on heat transfer in the BFB for CO2 methanation.
simulation.
− A 6. Conclusion
The axial 〈q 〉 for the 0% inlet N2 content is the highest among the
four inlet N2 contents because the temperature difference between the One of the key reactor designs for CO2 methanation is the effective
− A − A
bed and the wall outside is largest (see Fig. 11a). The 〈Tb − Twi 〉 for 0% heat removal from the catalyst bed because of the high amount of
inlet N2 content is also the highest within the fluidized-bed, as shown in exothermic reaction heat. The bubbling fluidized-bed (BFB) is suitable
− A − A for exothermic reactions because of its excellent gas–solid mixing and
Fig. 11b. However, the difference in the 〈Tb − Twi 〉 between the inlet N2 heat transfer. This study experimentally and numerically investigated
− A − A
contents is marginal compared to 〈q 〉. The 〈hc 〉 along h is uniform and is the hydrodynamics, reaction kinetics, and heat transfer of a bench-scale
the highest for the 0% inlet N2 content where the gas velocity is low and BFB reactor operated at 1 bar and a feed flow rate of 2 L/min with 77.5%
the solid holdup is high. The high solid concentration increases the N2 as the inert gas. A 3D gas–solid Eulerian computational fluid dy­
convective HTC (hc) over a certain range of ug/umf [16]. namics (CFD) model was coupled with a modified Syamlal–O’Brien drag
Table 5 reports the time- and volume-averaged physical properties of model and reaction kinetics for Ni-based catalysts. A grid convergence
the gas in the BFB for CO2 methanation. As the inlet N2 content de­ index (GCI) method was used to verify the numerical accuracy of the
creases from 77.5% to 0%, ρg and μg decrease because Tg increases. The CFD model. The CFD model was validated against experimental data,
specific heat capacity (cp,g) and thermal conductivity (κg) of gas increase such as pressure, temperature, and gas composition.
in general with the increase in temperature. CH4 has a higher cp,g and κg The pressure drop, solid velocity, solid volume fraction, temperature,

11
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

producer gas composition, and bed-to-wall heat transfer coefficient [7] Vandewalle J, Bruninx K, D’haeseleer W. Effects of large-scale power to gas
conversion on the power, gas and carbon sectors and their interactions. Energy
(HTC) were compared for four inlet N2 contents: 77.5% (experimental
Convers. Manage 2015;94:28–39. https://doi.org/10.1016/j.
condition), 50%, 25%, and 0% (no dilution). As the inlet N2 content enconman.2015.01.038.
decreased, the bed temperature increased due to exothermic CO2 [8] Castellani B, Rinaldi S, Morini E, Nastasi B, Rossi F. Flue gas treatment by power-
methanation, leading to thermal expansion, whereas the gas volume to-gas integration for methane and ammonia synthesis – Energy and environmental
analysis. Energy Convers Manage 2018;171:626–34. https://doi.org/10.1016/j.
decreased owing to the reduced mole number caused by the reaction. enconman.2018.06.025.
The gas volume reduction owing to the reaction for 0% inlet N2 content [9] Nguyen TDB, Seo MW, Lim Y-I, Song B-H, Kim S-D. CFD simulation with
influenced the hydrodynamics more significantly than the thermal experiments in a dual circulating fluidized bed gasifier. Comput Chem Eng 2012;
36:48–56. https://doi.org/10.1016/j.compchemeng.2011.07.005.
expansion. The time-averaged HTCs for the inlet N2 contents of 77.5%, [10] Kopyscinski J, Schildhauer TJ, Biollaz SMA. Employing catalyst fluidization to
50%, 25%, and 0% were 322, 338, 358, and 386 W/m2/K, respectively, enable carbon management in the synthetic natural gas production from biomass.
due to high solids concentration and high thermal conductivity of the Chem Eng Technol 2009;32:343–7. https://doi.org/10.1002/ceat.200800413.
[11] Witte J, Settino J, Biollaz SMA, Schildhauer TJ. Direct catalytic methanation of
gas. It was demonstrated that the bed temperature could be controlled biogas – Part I: New insights into biomethane production using rate-based
by feed dilution. An internal heat exchanger can be useful for the modelling and detailed process analysis. Energy Convers Manage 2018;171:
effective heat removal of BFBs for CO2 methanation. A polydispersed 750–68. https://doi.org/10.1016/j.enconman.2018.05.056.
[12] Chen JC, Grace JR, Golriz MR. Heat transfer in fluidized beds: design methods.
particle size would be more realistic than the monodispersed size used in Powder Technol 2005;150(2):123–32. https://doi.org/10.1016/j.
this study. powtec.2004.11.035.
[13] Kopyscinski J, Schildhauer TJ, Biollaz SMA. Methanation in a fluidized bed reactor
with high initial CO partial pressure: Part I-Experimental investigation of
CRediT authorship contribution statement
hydrodynamics, mass transfer effects, and carbon deposition. Chem Eng Sci 2011;
66(5):924–34. https://doi.org/10.1016/j.ces.2010.11.042.
Son Ich Ngo: Conceptualization, Data curation, Methodology, Soft­ [14] Vogtenhuber H, Pernsteiner D, Hofmann R. Experimental and numerical
investigations on heat transfer of bare tubes in a bubbling fluidized bed with
ware, Validation, Visualization, Writing - original draft, Writing - review
respect to better heat integration in temperature swing adsorption systems.
& editing, Funding acquisition. Young-Il Lim: Formal analysis, Super­ Energies 2019;12(14):2646. https://doi.org/10.3390/en12142646.
vision, Writing - review & editing, Funding acquisition, Project admin­ [15] Basu P. Circulating fluidized bed boilers: design, operation and maintenance.
istration. Doyeon Lee: Investigation, Resources, Funding acquisition. Cham, Switzerland: Springer; 2015. https://doi.org/10.1007/978-3-319-06173-3.
[16] Kim SW, Kim SD. Heat transfer characteristics in a pressurized fluidized bed of fine
Myung Won Seo: Supervision, Project administration. Sungwon Kim: particles with immersed horizontal tube bundle. Int J Heat Mass Transf 2013;64:
Investigation, Supervision. 269–77. https://doi.org/10.1016/j.ijheatmasstransfer.2013.04.045.
[17] Martin H. Heat transfer between gas fluidized beds of solid particles and the
surfaces of immersed heat exchanger elements, part I. Chem Eng Process 1984;18
Declaration of Competing Interest (3):157–69. https://doi.org/10.1016/0255-2701(84)80005-7.
[18] Nam H, Kim JH, Kim H, Kim MJ, Jeon S-G, Jin G-T, et al. CO2 methanation in a
bench-scale bubbling fluidized bed reactor using Ni-based catalyst and its
The authors declare that they have no known competing financial
exothermic heat transfer analysis. Energy 2021;214:118895. https://doi.org/
interests or personal relationships that could have appeared to influence 10.1016/j.energy.2020.118895.
the work reported in this paper. [19] Ngo SI, Nguyen TDB, Lim Y-I, Song B-H, Lee U-D, Choi Y-T, et al. Performance
evaluation for dual circulating fluidized-bed steam gasifier of biomass using quasi-
equilibrium three-stage gasification model. Appl Energ 2011;88(12):5208–20.
Acknowledgment https://doi.org/10.1016/j.apenergy.2011.07.046.
[20] Nguyen TDB, Ngo SI, Lim Y-I, Lee JW, Lee U-D, Song B-H. Three-stage steady-state
This research was supported by Basic Science Research Program model for biomass gasification in a dual circulating fluidized-bed. Energy Convers
Manage 2012;54(1):100–12. https://doi.org/10.1016/j.enconman.2011.09.019.
through the National Research Foundation of Korea (NRF) funded by the [21] Jia C, Dai Y, Yang Y, Chew JW. A fluidized-bed model for NiMgW-catalyzed CO2
Ministry of Education (grant number: 2020R1I1A1A01074184). This methanation. Particuology 2020;49:55–64. https://doi.org/10.1016/j.
work was also supported by the Korea Institute of Energy Technology partic.2019.05.004.
[22] Liu Y, Hinrichsen O. CFD simulation of hydrodynamics and methanation reactions
Evaluation and Planning (KETEP) and the Ministry of Trade, Industry in a fluidized-bed reactor for the production of synthetic natural gas. Ind Eng Chem
and Energy (MOTIE) of the Republic of Korea (No. 2019281010007B). Res 2014;53(22):9348–56. https://doi.org/10.1021/ie500774s.
[23] Sun L, Luo K, Fan J. Numerical investigation on methanation kinetic and flow
behavior in full-loop fluidized bed reactor. Fuel 2018;231:85–93. https://doi.org/
Supplementary data 10.1016/j.fuel.2018.05.042.
[24] Li J, Agarwal RK, Zhou L, Yang B. Investigation of a bubbling fluidized bed
Supplementary data to this article can be found online at https://doi. methanation reactor by using CFD-DEM and approximate image processing
method. Chem Eng Sci 2019;207:1107–20. https://doi.org/10.1016/j.
org/10.1016/j.enconman.2021.113863. ces.2019.07.016.
[25] Ngo SI, Lim Y-I. Multiscale Eulerian CFD of chemical processes: A review.
References ChemEngineering 2020;4:23. https://doi.org/10.3390/chemengineering4020023.
[26] Phuakpunk K, Chalermsinsuwan B, Putivisutisak S, Assabumrungrat S. Parametric
study of hydrogen production via sorption enhanced steam methane reforming in a
[1] Momeni M, Soltani M, Hosseinpour M, Nathwani J. A comprehensive analysis of a
circulating fluidized bed riser. Chem Eng Sci 2018;192:1041–57. https://doi.org/
power-to-gas energy storage unit utilizing captured carbon dioxide as a raw
10.1016/j.ces.2018.08.042.
material in a large-scale power plant. Energy Convers Manage 2021;227:113613.
[27] Ngo SI, Lim Y-I, Kim W, Seo DJ, Yoon WL. Computational fluid dynamics and
https://doi.org/10.1016/j.enconman.2020.113613.
experimental validation of a compact steam methane reformer for hydrogen
[2] Morgenthaler S, Ball C, Koj JC, Kuckshinrichs W, Witthaut D. Site-dependent
production from natural gas. Appl Energ 2019;236:340–53. https://doi.org/
levelized cost assessment for fully renewable Power-to-Methane systems. Energy
10.1016/j.apenergy.2018.11.075.
Convers Manage 2020;223:113150. https://doi.org/10.1016/j.
[28] Fattahi M, Hosseini SH, Ahmadi G, Parvareh A. Numerical simulation of heat
enconman.2020.113150.
transfer coefficient around different immersed bodies in a fluidized bed containing
[3] Ngo SI, Lim Y-I, Lee D, Go KS, Seo MW. Flow behaviors, reaction kinetics, and
Geldart B particles. Int J Heat Mass Transf 2019;141:353–66. https://doi.org/
optimal design of fixed- and fluidized-beds for CO2 methanation. Fuel 2020;275:
10.1016/j.ijheatmasstransfer.2019.06.058.
117886. https://doi.org/10.1016/j.fuel.2020.117886.
[29] Götz M, Lefebvre J, Mörs F, McDaniel Koch A, Graf F, Bajohr S, et al. Renewable
[4] Schaaf T, Grünig J, Schuster MR, Rothenfluh T, Orth A. Methanation of CO2 -
Power-to-Gas: A technological and economic review. Renew Energy 2016;85:
Storage of renewable energy in a gas distribution system. Energy Sustain Soc 2014;
1371–90. https://doi.org/10.1016/j.renene.2015.07.066.
4:2. https://doi.org/10.1186/s13705-014-0029-1.
[30] Ngo SI, Lim Y-I, Song B-H, Lee U-D, Lee J-W, Song J-H. Effects of fluidization
[5] Vu TT, Lim Y-I, Song D, Mun T-Y, Moon J-H, Sun D, et al. Techno-economic
velocity on solid stack volume in a bubbling fluidized-bed with nozzle-type
analysis of ultra-supercritical power plants using air- and oxy-combustion
distributor. Powder Technol 2015;275:188–98. https://doi.org/10.1016/j.
circulating fluidized bed with and without CO2 capture. Energy 2020;194:116855.
powtec.2015.02.017.
https://doi.org/10.1016/j.energy.2019.116855.
[31] Ngo SI, Lim Y-I, Song B-H, Lee U-D, Yang C-W, Choi Y-T, et al. Hydrodynamics of
[6] Qin Z, Zhao Y, Yi Q, Shi L, Li C, Yan X, et al. Methanation of coke oven gas over Ni-
cold-rig biomass gasifier using semi-dual fluidized-bed. Powder Technol 2013;234:
Ce/γ-Al2O3 catalyst using a tubular heat exchange reactor: Pilot-scale test and
97–106. https://doi.org/10.1016/j.powtec.2012.09.022.
process optimization. Energy Convers Manage 2020;204:112302. https://doi.org/
10.1016/j.enconman.2019.112302.

12
S. Ich Ngo et al. Energy Conversion and Management 233 (2021) 113863

[32] Niven RK. Physical insight into the Ergun and Wen and Yu equations for fluid flow [40] Xu J, Froment GF. Methane steam reforming, methanation and water-gas shift: I.
in packed and fluidised beds. Chem Eng Sci 2002;57:527–34. https://doi.org/ Intrinsic kinetics. AlChE J 1989;35(1):88–96.
10.1016/S0009-2509(01)00371-2. [41] Nguyen DD, Ngo SI, Lim Y-I, Kim W, Lee U-D, Seo D, et al. Optimal design of a
[33] Huilin Lu, Gidaspow D. Hydrodynamics of binary fluidization in a riser: CFD sleeve-type steam methane reforming reactor for hydrogen production from
simulation using two granular temperatures. Chem Eng Sci 2003;58(16):3777–92. natural gas. Int J Hydrog Energy 2019;44(3):1973–87. https://doi.org/10.1016/j.
https://doi.org/10.1016/S0009-2509(03)00238-0. ijhydene.2018.11.188.
[34] Gao J, Lan X, Fan Y, Chang J, Wang G, Lu C, et al. CFD modeling and validation of [42] Gaddis ES. VDI heat atlas. Heidelberg: Springer; 2010. https://doi.org/10.1007/
the turbulent fluidized bed of FCC particles. AIChE J 2009;55(7):1680–94. https:// 978-3-540-77877-6.
doi.org/10.1002/aic.v55:710.1002/aic.11824. [43] Boned C, Zéberg-Mikkelsen CK, Baylaucq A, Daugé P. High-pressure dynamic
[35] Gidaspow D. Multiphase flow and fluidization: Continuum and kinetic theory viscosity and density of two synthetic hydrocarbon mixtures representative of some
descriptions. San Diego: Academic Press Inc.; 1994. https://doi.org/10.1016/ heavy petroleum distillation cuts. Fluid Phase Equilib 2003;212(1-2):143–64.
C2009-0-21244-X. https://doi.org/10.1016/S0378-3812(03)00279-6.
[36] Syamlal M, O’Brien TJ. Computer simulation of bubbles in a fluidized bed. in: L.S. [44] Soomro M, Hughes R. The thermal conductivity of porous catalyst pellets. Can J
Fan, (Ed.). Fluidization and Fluid Particle Systems: Fundamentals and Chem Eng 1979;57(1):24–8. https://doi.org/10.1002/cjce:v57:110.1002/cjce:
Applications, Michigan: AIChE Symposium Series No. 270; 1989. pp. 22-31. 5450570104.
[37] Schaeffer DG. Instability in the evolution equations describing incompressible [45] Tran BV, Ngo SI, Lim Y-I, Bae K, Lee DH, Go K-S, et al. Hydrodynamics of air-
granular flow. J Differ Equ 1987;66(1):19–50. https://doi.org/10.1016/0022- kerosene bubble column under elevated pressure in homogeneous flow regime.
0396(87)90038-6. Chin J Chem Eng 2020. https://doi.org/10.1016/j.cjche.2020.08.031 (in press).
[38] Gunn DJ. Transfer of heat or mass to particles in fixed and fluidised beds. Int J Heat [46] Roache PJ. Verification of codes and calculations. AIAA J 1998;36(5):696–702.
Mass Transf 1978;21(4):467–76. https://doi.org/10.1016/0017-9310(78)90080-7. https://doi.org/10.2514/2.457.
[39] Zimmermann S, Taghipour F. CFD modeling of the hydrodynamics and reaction [47] Molerus O, Burschka A, Dietz S. Particle migration at solid surfaces and heat
kinetics of FCC fluidized-bed reactors. Ind Eng Chem Res 2005;44:9818–27. transfer in bubbling fluidized beds-I. Particle migration measurement systems.
https://doi.org/10.1021/ie050490+. Chem Eng Sci 1995;50(5):871–7. https://doi.org/10.1016/0009-2509(94)00445-
W.

13

You might also like