You are on page 1of 9

Minerals Engineering 149 (2020) 106238

Contents lists available at ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

CFD modeling and simulation of gas-phase extraction processes in fluidized T


bed reactor

Le Xie, Jundong Zhu, Jiang Hu, Chongwen Jiang
College of Chemistry and Chemical Engineering, and Hunan Provincial Key Laboratory of Efficient and Clean Utilization of Manganese Resources, Central South University,
Changsha 410083, Hunan, China

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

Keywords: Recently, the gas-phase organic ligand extraction of metals from low-grade sources using the fluidized bed
Computational fluid dynamics (CFD) reactor (FBR) technique has garnered significant attention owing to its simplicity and advantages in reducing
Binary-mixture particles energy costs. In this study, a computational fluid dynamics (CFD) coupled model including a three-phase
Minerals processing Eulerian model, kinetic theory of granular flow model, and reaction kinetics model was proposed to investigate
Extractive reaction
gas-phase extraction processes encountered in an FBR. The developed three-phase CFD coupled model was first
Fluidized bed reactor (FBR)
validated by comparing the obtained simulation results with the experimental data of cumulative iron extraction
at different reaction temperatures and the classical calculated data of the pressure drop. Subsequently, the effects
of the inlet gas velocity and particle phase holdup on the extraction were studied. Accordingly, both particle
fluidization behaviors and extractive reaction kinetics characteristics were obtained. The simulation results
indicated that the mentioned operating conditions significantly affected the distribution of particle phase holdup
and subsequently influenced the cumulative iron yield through an extractive reaction. Furthermore, the de-
veloped CFD coupled model is beneficial for optimizing gas-phase extractive reaction process in FBRs.

1. Introduction extractants that pass through a heated feed material and react selec-
tively with metal oxides (Allimann-Lecourt and Bailey, 2002; Dyk et al.,
A large number of slags serving as by-products in metallurgy pro- 2010; Potgieter et al., 2006). Some experimental studies have been
cesses are produced annually worldwide (Liu et al., 2016; Piatak et al., performed on the effect of extractants on extraction efficiency
2015; Sarfo et al., 2017; Shen and Forssberg, 2003). In general, these (Allimann-Lecourt and Bailey, 2002; Sun and Waters, 2014) and gas-
slags should be handled with the utmost care because they contain phase extraction reaction kinetics behaviors (Dyk et al., 2010; Potgieter
harmful or heavy metals that can cause environmental problems. Fur- et al., 2006). Many detailed experimental investigations of the gas-
thermore, most slags that contain variable amounts of valuable metals phase extraction of metals from oxides have been reported.
are in fact a secondary resource of metals. Therefore, it is crucial to Generally, gas-phase extraction reactions are performed in fluidized
recover metals from these low-grade sources to recycle and reuse re- bed reactors (FBRs) owing to their excellent mass and heat transfer
sources, as well as for environmental protection (Piatak et al., 2015; performance. For gas-phase extraction in an FBR, oxide particles typi-
Shen and Forssberg, 2003). cally coexist with sand particles. When the gas velocity is higher than
In the last several decades, many types of conventional technolo- the minimum fluidization velocity, the particle mixture starts to flow
gies, such as hydrometallurgy (Arslan and Arslan, 2002; Banza et al., (fluidization). Clearly, a gas-phase extraction system is a typical binary-
2002; Petersen, 2016), pyrometallurgy (Gyurov et al., 2011; Sarfo et al., mixture particle system (gas–solid-solid three-phase) in which the
2017; Ye et al., 2003; Zhang et al., 2007), and bioleaching (Akcil et al., particle–particle mixing quality is critical in determining the gas-phase
2007; Carranza et al., 2009; Potysz et al., 2018), have been proposed to extraction reaction rate. Therefore, it is crucial to comprehensively
recover metals from slags. Such studies have confirmed that a higher understand the relationship between fluid flow and extraction reactions
metal recovery rate can be realized, and the resulting data are useful for to optimize this process, which is characterized by multiphase flow and
the optimization of the recovery process. In addition, gas-phase ex- multiscale characteristics. As a supplement, computational fluid dy-
traction has emerged as a powerful technique, and it is employed to namics (CFD) simulations can provide flow field information that can
recover metals from their oxides. Volatile organic compounds serve as explain the experimental observations. Additionally, both the evolution


Corresponding author.
E-mail address: jcwcsu@csu.edu.cn (C. Jiang).

https://doi.org/10.1016/j.mineng.2020.106238
Received 3 June 2019; Received in revised form 20 January 2020; Accepted 30 January 2020
Available online 07 February 2020
0892-6875/ © 2020 Elsevier Ltd. All rights reserved.
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Nomenclature t Time, (s)


T Temperature, (K)
CD Drag coefficient, (-) U Slip velocity, (m s−1)
Cfr Friction coefficient, (-) v Velocity, (m s−1)
d Diameter of particles (m) w Mass fraction (-)
D Diffusion coefficient (m2 s−1) y Yield (-)
e Restitution coefficient, (-)
F Interphase force (N m-3) Greek letters
g Gravitational acceleration (m s−2)
g0 Radial distribution function, (-) α Volume fraction, (-)
h Specific enthalpy, (kJ kg−1 K−1) ζ Inter-phase drag force coefficient, (kg m−3 s−1)
I Unit tensor, (-) β Inter-phase drag force coefficient, (kg m−3 s−1)
k reaction rate constant (s−1) γs Collisional dissipation energy, (kg m−1 s−3)
kΘ KTGF model parameter, (-) ρ Density, (kg m−3)
mi Gas-solid drag model parameter, (-) τ̄¯ Stress tensor, (Pa)
msp Interphase mass transfer, (kg m−3 s−1) μ Viscosity, (Pa s)
ni Gas-solid drag model parameter, (-) Θ Granular temperature, (m2 s−2)
Nc Courant number, (-) λ Solid bulk viscosity, (Pa s)
p Pressure, (Pa) θ Angle of internal friction, (deg)
Re Reynolds number, (-) κ Thermal conductivity, (W m−1 K−1)
Si The source term of species i (kg m−3 s−1)

of fluidization and the gas-phase extraction reaction in an FBR can be model was proposed for the first time to study the multiphase flow
visualized by a CFD simulation. reaction behaviors of the gas-phase extraction of iron from its oxide in
Gas–solid hydrodynamic characteristics in FBRs can be numerically an FBR. The particle fluidization characteristics and gas-phase ex-
investigated by many multiphase models, which include the multi fluid tractive reaction kinetics behaviors were investigated. The developed
model (MFM), dense discrete particle model (DDPM) and CFD-discrete model was validated by the experimental data of cumulative iron ex-
element model (CFD-DEM). A deep understanding of the advantages traction and the classical calculated data of pressure drops.
and disadvantages of these models is inherently required for reliable Subsequently, the CFD coupled model was employed to explore the
simulation results. In recent years, a considerable effort has been de- effect of inlet gas velocity and particle phase holdup. The hydro-
voted to evaluating these different FBR modeling approaches (Chen and dynamic characteristics of the gas–solid-solid three-phase flow reaction
Wang, 2014; Le Lee and Lim, 2017; Nikolopoulos et al., 2017; behaviors in an FBR were obtained.
Ostermeier et al., 2019a; 2019b; 2019c; Stroh et al., 2018; Wang et al.,
2013). The MFM and DDPM approaches require fewer computing re-
sources and have the potential for industrial applications. The CFD- 2. Mathematical modeling and simulation details
DEM approach can give more details about flow patterns. However, the
as-mentioned studies focused on gas–solid two-phase flow instead of To present a comprehensive description of the process of gas-phase
binary-mixture particle systems, which have received extensive atten- extraction in an FBR, a multiple model coupling approach is typically
tion in recent years (Benyahia, 2008; Chao et al., 2011; Das et al., 2008; required. In this study, the developed coupled model is composed of a
Jia et al., 2012; Mathiesen et al., 2000). Based on the kinetic theory of three-fluid CFD model (gas phase, oxide particle phase, and sand par-
granular flow (KTGF), Lu’s group proposed some constitutive relations ticle phase), a KTGF model, and a gas-phase extraction reaction kinetics
to calculate particle phase viscosity (Lu et al., 2000; 2001). The avail- model. As the three-fluid CFD and KTGF models are widely applied in
ability and suitability of the developed models were validated by ex- simulating binary-mixture particle systems, herein, they are summar-
perimental data from a riser (Lu and Gidaspow, 2003) and a bubbling ized in Table 1.
fluidized bed (Sun et al., 2005). In addition, other CFD simulations As shown, each phase (three phases in total) has a set of governing
focused on investigating the effects of numerical methods (i.e., DEM equations, which include the continuity equation, momentum equation,
and the Eulerian approach) (Cheng et al., 2014; Jalali and Hyppänen, energy equation, and species equation. For the solid phase, the extra
2010) and gas–solid/solid–solid interactions (Rong et al., 2014; Zhang KTGF model is required to calculate solid phase properties, such as solid
et al., 2017; Zhou and Wang, 2015) on the fluidization behaviors of pressure, solid shear viscosity, and solid bulk viscosity. For the species
binary-mixture particles. These simulation results presented a com- equation, there are two independent equations in the gas phase.
prehensive description of the distribution of solid particles in an FBR. Additionally, the volume fraction of the oxide particles should be cal-
More recently, CFD models have been developed and employed to in- culated. The interphase mass transfer (msp) can be calculated based on
vestigate the flotation process (Wang et al., 2018), including predicting the gas phase or solid phase. In this study, obviously, the interphase
the flotation rate constant (Karimi et al., 2014a; 2014b) and de- mass transfer rate from the oxide particle phase to the gas phase just
termining the gas dispersion characteristics and particle–bubble inter- equals the consumption rate of Fe2O3 particles, and it is integrated into
actions in flotation cells (Basavarajappa and Miskovic, 2016; the CFD model using user-defined functions.
Hassanzadeh et al., 2018). The above studies provided a comprehensive It should be noted here that, for the interphase force, only the drag
investigation of the fluidization behaviors of gas–solid binary-mixture force is included in the momentum equation, and the proposed CFD
particles in FBRs. However, the relationship between the particle flui- coupled model may be only suitable for a laboratory scale FBR. When it
dization and the gas–solid reaction is still unclear because chemical is extended to industrial scale reactors, the drag model must be replaced
reactions are mostly ignored. In particular, the literature pertaining to by the filtered drag correlations recommended by Sundaresan's group
the gas-phase extraction of metals from oxides in FBR based on CFD (Milioli et al., 2013; Igci and Sundaresan, 2011). Additionally, an ap-
techniques is limited. parent extraction reaction kinetics model was used for simplicity.
In this study, a CFD-KTGF model coupled with a reaction kinetics Therefore, an accurate reaction kinetics model based on the reaction
mechanism is of great importance for improving the accuracy of

2
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Table 1
Governing equations of the three-liquid binary-mixture particle system.
Models Equations

Continuity equation
+ ∇·(αq ρq→

(α ρ ) vq) = msp, q = s or g
∂t q q
2
αs + αg = ∑i = 1 αsi + αg = 1
Momentum equation →→ → →
(α ρ → + ∇·(αq ρq v v ) = −αq ∇p − ∇·τ¯q − msp v − F + αq ρq→

v) g
∂t q q q q q q

τ¯g = αg μg [∇→
vg + (∇→
vg )T ] − αg μg ∇→
2
vg I
3

τ¯si = αsi μsi [∇→
vsi + (∇→
vsi )T ] + (λ si − μsi ) αsi ∇→
2
vsi I
3
Energy equation ∂ ∂pq →
(α ρ h ) + ∇·(αq ρq vq hq) = −αq + τ¯q: ∇ v − ∇·qq
∂t q q q ∂t q
n
+ ∑p = 1 (msp hsp − mps hps ), qq = −αq κ q ∇Tq
Species equation ∂
(α ρ w ) + ∇·(αq ρq vq wi ) = ∇·(Di ∇ (ρq wi )) + Si
∂t q q i
Granular temperature equation
[ (ρ α Θ ) + ∇·(ρsi αsi → ∇→
3 ∂
vsi Θsi )] = (−psi I¯ + τ¯si ): vsi + ∇·(k Θsi ∇Θsi )
2 ∂t si si si
− γΘsi − 3(βg − si + ζs, ij )Θsi
150ρsi dsi Θsi π 6 Θsi
k Θsi = [1 + αsi g0, ij (1 + es, ij )]2 + 2αsi2 ρsi dsi g0, ij (1 + es, ij )
384(1 + es, ij ) g0, ij 5 π
12(1 − es2, ij ) g0, ij
γΘs = ρsi αsi2 Θ1.5
si
dsi π
Solid pressure 2 dsi + dsj
Psi = αsi ρsi Θsi + 2αsi ρsi Θsi ∑ j = 1 [ ]3 (1 + es, ij ) g0, ij αsj
2dsi
gsi dsi + gsj dsj αs 3 2 α sj
g0, ij = , gsi = [1 − ( )1/3]−1 + dsi ∑ j = 1
dsi + dsj α s,max 2 dsj
α si,max
if mi ⩽
α si,max + (1 − α si,max ) α sj,max

dsj
αs,max = [αsi,max − αsj,max + (1 − )(1 − αsi,max ) αsj,max ]
dsi
mi
× [αsi,max + (1 − αsi,max ) αsj,max ] + αsj,max
α si,max
Packing limit dsj
else αs,max = (1 − )[αsi,max + (1 − αsi,max ) αsj,max ](1 − mi) + αsi,max
dsi
Solid bulk viscosity 4 Θsi
λ si = α ρ d g (1 + es, ij )
3 si si si 0, ij π
μsi = μsi, col + μsi, kin + μsi, fr
4 Θsi 0.5
μsi, col = αsi ρsi dsi go, ij (1 + es, ij )( )
5 π
α si dsi ρsi Θsi π 2
μs, kin = [1 + (1 + es, ij )(3es, ij − 1) αsi g0, ij ]
6(3 − es, ij ) 5
Solid shear viscosity psi sin θsi
μsi, fr = .
2 I

F = βg − si (→
vsi − →
Re −f (Re, αg )
vg ), βd = C α
24 d0 g
βg − si 0.5αg
= + 0.5(1 − αg ) ni2 + 0.5ni
βd ∑i2= 1 (mi / ni2)

f (Re,αg ) = 2.65(1 + αg ) − (5.3 − 3.5αg ) αg2 exp[−0.5(1.5 − log Re)2]


Gas-solid drag force model (Rong et al., 2014) αg ρg < d > Uslip 2 α si
Re = , < d > = [ ∑i = 1 (mi / di )]−1 , ni = di/ < d > , mi =
μg αs

Solid-solid drag force model (Syamlal, 1987) π π2


3(1 + es, ij )( + Cfr , ij ) α si ρsi α sj ρsj (dsi + dsj )2g0, ij
ζs, ij = 2 8
3 + ρ d3 ) |→
vsi − →
vsj|
2π (ρsi dsi sj sj

predictions. − ln(1 − y ) = kt (1)

where y is the extraction rate of metallic Fe, and k is the reaction rate
2.1. Reaction kinetics model constant which is a function of temperature. In this study, the reported
experimental data were employed to determine the apparent reaction
At high temperatures, iron(III) oxide (Fe2O3) reacts readily with rate constant based on the least-squares method.
acetylacetone; the stoichiometric equation is described as follows: As shown in Fig. 1, high correlation coefficients (> 0.92) for the
Fe2 O3 (s ) + 6H(C5 H7 O2) (g ) = 2Fe(C5 H7 O2)3 (g ) + 3H2 O (g ) extraction reactions were obtained when the experimental data were
fitted to a first-order rate equation. The slope of the line is the reaction
This gas–solid reaction can be characterized by a noncatalytic rate constant that is employed to determine the apparent activation
shrinking particle model, and no product layer is formed because the energy. The linear fitting results are shown in Fig. 2. The apparent
products are gasified briefly under high-temperature conditions. In activation energy is 62.91 kJ/mol. It is worth noting that the apparent
general, the gas-phase extraction rate is determined by a chemical re- reaction rate constant is dependent on the initial Fe2O3 charge (see
action, mass transport, or both. Van Dyk et al. (2010) investigated the Fig. 1B). In this study, the determined reaction kinetics model should be
effect of temperature, metal oxide concentration, and ligand flow rate integrated into a CFD model to describe the reaction kinetics char-
on extraction; they concluded that the gas-phase extraction reaction acteristics.
obeys first-order kinetics the best. For an irreversible first-order reac-
tion, the following well-known equation can be used:

3
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Fig. 1. Results of the linear regression for cumulative iron extraction and reaction time under different operating conditions: (A) reaction temperature; (B) Fe2O3
charge.

2.3. Simulation details

All simulations were performed in a two-dimensional (2D) FBR with


a diameter of 15 mm and a height of 500 mm. Acetylacetone was one of
the reactants and also served as a fluidizing agent. A sand particle has
an average size of approximately 64 μm, which is similar to that of an
Fe2O3 particle. However, these two particles are different in density
and, thus, are considered as two different solid phases. The two solid
phases had a total mass of 50 g, and the initial volume fraction of the
solid phase was 0.6. Therefore, when the Fe2O3 charge was 1 wt%, the
initial bed height was 0.2128 m. It should be noted that the volume
fraction of the solid phase was very low in this study. Therefore, we did
not evaluate the effect of frictional force on the mixing of solid parti-
cles. A constant solid–solid friction coefficient (Cfr =0.15) and a default
frictional viscosity model (Schaeffer model) were used in the CFD si-
mulation. The detailed settings, initial and boundary conditions and
model parameters are listed in Table 2.
The FBR was created and meshed by structured quadrilateral cells
Fig. 2. Results of the linear regression for apparent reaction rate constant and using the commercial software GAMBIT 2.3.16 (Ansys Inc.). The cell
reaction temperature. size was 0.0005 m. Consequently, 30,000 cells were selected for all the
CFD simulations. For binary-mixture particle systems, Coroneo et al.
(2011) recommended that a reasonable Courant number (Nc ) (see Eq.
2.2. Model coupling mechanism
(2)) should be in the range of 0.028–0.15. In this study, the gas velo-
city was 0.01–0.06 m/s, and the time step was 0.001 s. Therefore, the
Fig. 3 displays the coupling scheme of the proposed submodels. At
the beginning of the simulation, the velocity, pressure, phase holdup, Courant number was 0.02–0.12.
temperature, and species mass fraction in the FBR are initialized. The Δt
KTGF model is solved to obtain the solid phase pressure and viscosity. Nc = vg
Δcell (2)
The gas–solid drag force is calculated based on the Rong model (Rong
et al., 2014) recommended by Zhang et al. (2017). Subsequently, In addition, the developed 2D CFD coupled model was solved by the
continuity and momentum equations can be solved to obtain the ve- commercial code FLUENT (Ansys Inc.). The governing equations were
locity field information in each cell. Next, the temperature and species discretized by the QUICK scheme and solved under unsteady condi-
mass fraction are transferred into the reaction kinetics model to cal- tions. The pressure-based solver with an absolute velocity formulation
culate the reaction rate, including the consumption rate of reactants was used. The gradient was computed by a Least Squares Cell Based
and the formation rate of products, which are the source terms of the scheme. The pressure–velocity coupling was realized by the SIMPLE
species transport equations. Additionally, these source terms should be scheme. In this study, the fixed time stepping method was used
transferred into the CFD three-phase flow model (interphase mass (Δt = 0.001 s), and the convergence criterion was set at 1 × 10−6. All
transfer). Subsequently, the species mass fraction distributions are up- CFD simulations were executed on a 2.2-GHz Intel 2 Central Processing
dated by solving these models. In this study, the coupled drag force Unit (16 cores) with a 128 GB RAM. When the CFD Euler model, the
model and reaction kinetics model were written in C and were in- KTGF model, and reaction kinetics model were solved together, about
tegrated into a CFD model using user-defined functions. The above two weeks were needed to simulate the physical time of 120 min.
mentioned iterative process was performed repeatedly until a specific
criterion of physical time was satisfied. For more details about the
model coupling mechanism, please refer to Fig. 3.

4
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Fig. 3. Solution procedure for the CFD coupled model.

Table 2
Boundary conditions and model parameters.
Descriptions Values

Inlet boundary condition Velocity inlet


Outlet boundary condition Pressure outlet
Wall boundary condition No slip; constant temperature
Initial bed height 0.2016–0.2128 m
Initial volume fraction of solid phase 0.6
Reaction temperature 463–523 K
Inlet gas velocity 0.01–0.06 m/s
Solid-solid friction coefficient 0.15
Column diameter 15 mm
Column height 500 mm
Cell size 0. 5 mm
Convergence criteria 1 × 10−6
Time step 0.001 s

Fig. 5. Comparisons between simulation results (CFD) and experimental data


(Exp.) for the pressure drop.

3. Results and discussion

3.1. Model validation

3.1.1. Kinetics model validation


To study the effect of reaction temperature on the gas-phase ex-
traction reaction kinetics behavior, experiments have been performed
in a FBR by Van Dyk et al. (2010). The experimental conditions used
were as follows: acetylacetone flow rate = 1 ml/min, Fe2O3
charge = 1 wt% (total mass of binary-mixture particles = 50 g), and
T = 190, 210, 230, 250 °C. In this study, CFD simulations were per-
formed under identical operating conditions and FBR structure para-
meters to compare the simulation results with the experimental data in
terms of cumulative iron extraction. It is worth noting that the inlet gas
Fig. 4. Comparisons between simulation results (CFD) and experimental data velocity was determined by the ideal gas state equation. Fig. 4 displays
(Exp.) for cumulative iron extraction under four different reaction tempera- the predicted cumulative iron extraction profile along with the reaction
tures. time.
From Fig. 4, the cumulative iron extraction is highly dependent on

5
L. Xie, et al. Minerals Engineering 149 (2020) 106238

was compared with that obtained by an empirical equation (Eq. (3)) to


validate the proposed CFD coupled model. The comparison results are
shown in Fig. 5. Additionally, the time evolution of particle fluidization
is presented in Fig. 6. As shown, a steady fluidization can be developed
easily at the inlet gas velocity of 0.035 m/s. In a steady fluidization, the
bed height is ~0.321 m, and the calculated bed pressure drop is
2766 Pa. As shown in Fig. 5, the simulated results are in good agree-
ment with the calculated data, thus validating the CFD coupled model.

3.2. Model application

In this section, the effects of some key operating parameters (i.e.,


inlet gas velocity and particle phase holdup) on the particle fluidization
characteristics and reaction kinetics behaviors are investigated based
on the validated model.

3.2.1. Effects of the inlet gas velocity


The gas–solid drag force is an important interphase force de-
termining particle fluidization behaviors that can subsequently and
significantly affect the distributions of reaction rates and products.
Therefore, the inlet gas velocity should be addressed carefully ac-
cording to the established gas–solid drag force model (see Table 1). In
this section, the effect of the inlet gas velocity (0.01, 0.02, 0.04, and
0.06 m/s) on the results of the simulation cases is discussed. The time
evolutions of particle fluidization are obtained and shown in Fig. 7.
In general, with an increase in the inlet gas velocity, the particle bed
Fig. 6. Time evolution of the volume fraction of the solid phase at the inlet gas will pass through the fixed bed, critical fluidization, particulate ex-
velocity of 0.035 m/s. pansion, bubbling fluidization, and fast fluidization, and so on. As
shown, the bed expansion height increases from 0.25 to 0.378 m with
the reaction temperature. With an increase in reaction temperature an increase in inlet gas velocity from 0.01 to 0.06 m/s. From Fig. 7A
from 463 K to 523 K, the cumulative iron extraction increases from 12% and 7B, when the inlet gas velocity is lower than 0.02 m/s, the particle
to 65% after 120 min. Therefore, FBR technology has similar ad- bed is in the state of particulate expansion, in which no bubbles are
vantages as other technologies when it comes to process efficiency. formed and the particle bed is uniform. As the inlet gas velocity is in-
Generally, advantages and disadvantages coexist for FBR technology. creased to 0.06 m/s, however, bubbling fluidization is observed (see
Although FBR has excellent mass and heat transfer performance and is Fig. 7D). The bubble size increases gradually as it rises along the bed.
convenient in operation, the solid phase has a considerable range of The fluidized bed reaches a steady fluidized state after ~4 s. For
residence time distribution, leading to polydispersed product distribu- binary-mixture particles, large and heavy particles tend to reside at the
tion and low production yield; the operating air velocity should be bed bottom, while small and light particles are more easily con-
chosen carefully to ensure stable fluidization. Furthermore, complex centrated on the bed surface (Zhou and Wang, 2015). In this study, the
fluid mechanics and transport phenomena hinder the optimized design oxide particles and sand particles are of a similar size but with different
and scale-up of FBRs. Therefore, it is of great importance to perform a densities. However, the segregation phenomenon of the binary-mixture
comprehensive investigation of the gas-phase extraction processes in particles was not observed in this study. The extremely low levels of
FBRs based on CFD technology. oxide particles and the same size of oxide particles and sand particles
Additionally, the simulation results revealed the same trend as the may be the possible reasons. Thus, a homogeneous mixing of the oxide
experimental data. Potgieter et al. (2006) concluded that the gas-phase particle phase and sand particle phase was obtained.
extraction reaction corresponded to a first-order reaction, which was Fig. 8 displays the time evolution of the bed expansion height. In the
derived based on the hypothesis that perfect mixing occurs. In this particulate expansion state, the bed expansion height increases uni-
study, a first-order reaction kinetics model was integrated into a CFD formly with flow time and finally stabilizes at a constant value. In the
model to describe the transport/dynamics of bubbles, particle fluidi- bubbling fluidization state, the bed expansion height fluctuates owing
zation, and reaction kinetics behaviors. In the CFD simulation, how- to the formation and movement of bubbles.
ever, the distribution of Fe2O3 particles is instantaneous. As a result, the
distribution of the extraction rate changes spatially and temporally as 3.2.2. Effect of particle phase holdup
the reaction progresses. These differences are finally reflected in the According to the developed CFD coupled model, particle phase
cumulative iron extraction. Therefore, an accurate reaction kinetics holdup is another important parameter determining the particle flui-
model based on a reaction mechanism is inherently required to improve dization behaviors and reaction kinetics characteristics. In this section,
the reliability of CFD simulation results. three Fe2O3 charges (i.e., 1%, 3%, and 10%) were selected from the
study by Van Dyk et al. (2010). For binary-mixture particle systems, it is
widely known that particle fluidization occurs with increased difficulty
3.1.2. Hydrodynamics characteristics validation with an increase in particle-phase holdup.
For binary-mixture particles in an FBR, bed pressure drop is an As shown in Fig. 9, when the Fe2O3 charge is 1%, the steady bed
important parameter in evaluating hydrodynamic performance. expansion height is ~0.321 m. When the Fe2O3 charge increased to
Empirical equations are available to calculate the bed pressure drop 10%, however, the steady bed expansion height was only 0.278 m. It is
(Romeo et al., 2011; Singh et al., 2008): noteworthy that the initial bed height cannot be ignored under different
ΔP = (1 − αg ) hmf (ρs − ρg ) g Fe2O3 charges. The initial bed height declines by 0.011 m when the
(3)
Fe2O3 charge increases from 1% to 10%. The density of the oxide
In this study, the bed pressure drop calculated by the CFD model particles is higher than that of the sand particles; thus, defluidization is

6
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Fig. 7. Time evolutions of the volume fraction of the solid phase under different inlet gas velocities: (A) 0.01 m/s; (B) 0.02 m/s; (C) 0.04 m/s; and (D) 0.06 m/s.

Fig. 8. Time evolution of the bed expansion height under different inlet gas Fig. 9. Time evolution of the bed expansion height under three different Fe2O3
velocities. charges.

7
L. Xie, et al. Minerals Engineering 149 (2020) 106238

Fig. 11 displays the contour plots of the reaction rate in an FBR. As


shown, the distribution of the gas-phase extraction rate is highly de-
pendent on the distribution of the phase holdup of the Fe2O3 particles.
After the Fe2O3 charge increased from 1% to 3% and 10%, the average
reaction rates (the consumption rate of Fe2O3) were 0.000749,
0.002038, and 0.002526 kg/m3/s, respectively. Thus, the actual
amount of iron being extracted increased. Van Dyk et al. (2010) ob-
served that when the Fe2O3 charge was 10 wt%, the iron that was ex-
tracted was three times that obtained from a 1 wt% charge after 4 h.

4. Conclusions

A CFD coupled model involving a three-phase Eulerian model, KTGF


model, and reaction kinetics model was proposed to investigate gas-
phase extraction processes occurring in an FBR. The developed CFD
coupled model was validated by experimental data in terms of the cu-
mulative iron extraction at different reaction temperatures and the
classical calculated data of the pressure drop. Based on the validated
Fig. 10. Comparisons between simulation results (CFD) and experimental data model, the effects of inlet gas velocity and particle phase holdup on the
(Exp.) for cumulative iron extraction under three different Fe2O3 charges. extraction process were investigated. As the inlet gas velocity increased
from 0.01 to 0.06 m/s, the particles bed underwent particulate ex-
observed with an increase in oxide particle phase holdup. pansion into bubbling fluidization and the bed expansion height in-
Fig. 10 shows the time evolution of the cumulative iron extraction creased from 0.25 to 0.378 m. Accordingly, both the particle fluidiza-
under three different Fe2O3 charges. The CFD simulation conditions tion behaviors and extractive reaction kinetics characteristics were
were identical to the reported experimental conditions: the inlet gas obtained under three different Fe2O3 charges. Defluidization was ob-
velocity was 0.0345 m/s, the total mass of the solid particles was 50 g, served with an increase in oxide particle phase holdup. The simulation
and the reaction temperature was 250 °C. The CFD predictions are in results demonstrated that the distribution of the gas-phase extraction
good agreement with the experimental data. As discussed above, the rate was highly dependent on the distribution of Fe2O3 particle phase
gas-phase extraction reaction obeys the first-order kinetics best in terms holdup, and the cumulative iron extraction decreased with an increase
of the Fe2O3 charge under different reaction temperatures. Ad- in Fe2O3 charges. Meanwhile, the average reaction rate and actual
ditionally, these reaction kinetics characteristics are observed under amount of iron being extracted increased.
different Fe2O3 charges, while the apparent reaction rate constant de-
pends on the initial Fe2O3 charge. As shown in Fig. 10, the cumulative CRediT authorship contribution statement
iron extraction decreases with an increase in Fe2O3 charge. For an ir-
reversible first-order reaction, the extraction rate is independent of the Le Xie: Conceptualization, Methodology, Software, Writing - ori-
initial concentration of the reactant. Herein, the decreased reaction rate ginal draft. Jundong Zhu: Writing - review & editing. Jiang Hu:
with the increase in the initial Fe2O3 charge may indicate that the gas- Writing - review & editing. Chongwen Jiang: Supervision, Project
phase extraction reaction is reversible under high temperatures. administration, Funding acquisition.
Therefore, it is necessary to develop a comprehensive mass transfer-
reaction kinetics model to gain a deep understanding of the relationship Declaration of Competing Interest
between particle fluidization behaviors and reaction kinetics char-
acteristics. The authors declare that they have no known competing financial

Fig. 11. Contour plots of the reaction rate (kg/m3/s) in the FBR under different Fe2O3 charges: (A) 3% Fe2O3; (B) 10% Fe2O3.

8
L. Xie, et al. Minerals Engineering 149 (2020) 106238

interests or personal relationships that could have appeared to influ- Lu, H., Liu, W., Bie, R., Yang, L., Gidaspow, D., 2000. Kinetic theory of fluidized binary
ence the work reported in this paper. granular mixtures with unequal granular temperature. Phys. A 284, 265–276.
Mathiesen, V., Solberg, T., Hjertager, B.H., 2000. An experimental and computational
study of multiphase flow behavior in a circulating fluidized bed. Int. J. Multiphas.
Acknowledgements Flow. 26, 387–419.
Milioli, C.C., Milioli, F.E., Holloway, W., Agrawal, K., Sundaresan, S., 2013. Filtered two-
fluid models of fluidized gas-particle flows: new constitutive relations. AIChE J. 59,
The authors thank the National Natural Science Foundation of 3265–3275.
China (No. 21476267) and Natural Science Foundation of Hunan Nikolopoulos, A., Stroh, A., Zeneli, M., Alobaid, F., Nikolopoulos, N., Ströhle, J., Karellas,
Province (No. 2018JJ2482) and the Center for High Performance S., Epple, B., Grammelis, P., 2017. Numerical investigation and comparison of coarse
grain CFD – DEM and TFM in the case of a 1MWth fluidized bed carbonator simu-
Computing, Shanghai Jiao Tong University for supporting this work. lation. Chem. Eng. Sci. 163, 189–205.
Ostermeier, P., DeYoung, S., Vandersickel, A., Gleis, S., Spliethoff, H., 2019a.
References Comprehensive investigation and comparison of TFM, DenseDPM and CFD-DEM for
dense fluidized beds. Chem. Eng. Sci. 196, 291–309.
Ostermeier, P., Fischer, F., Fendt, S., DeYoung, S., Spliethoff, H., 2019b. Coarse-grained
Akcil, A., Ciftci, H., Deveci, H., 2007. Role and contribution of pure and mixed cultures of CFD-DEM simulation of biomass gasification in a fluidized bed reactor. Fuel 255,
mesophiles in bioleaching of a pyritic chalcopyrite concentrate. Miner. Eng. 20, 115790.
310–318. Ostermeier, P., Vandersickel, A., Gleis, S., Spliethoff, H., 2019c. Numerical approaches for
Allimann-Lecourt, C., Bailey, T.H., 2002. Purification of combustion fly ashes using the modeling gas–solid fluidized bed reactors: comparison of models and application to
SERVO process. J. Chem. Technol. Biot. 77, 260–266. different technical problems. J. Energ. Resour-ASME. 141, 070707.
Arslan, C., Arslan, F., 2002. Recovery of copper, cobalt, and zinc from copper smelter and Petersen, J., 2016. Heap leaching as a key technology for recovery of values from low-
converter slags. Hydrometallurgy 67, 1–7. grade ores – A brief overview. Hydrometallurgy 165, 206–212.
Banza, A.N., Gock, E., Kongolo, K., 2002. Base metals recovery from copper smelter slag Piatak, N.M., Parsons, M.B., Seal II, R.R., 2015. Characteristics and environmental aspects
by oxidising leaching and solvent extraction. Hydrometallurgy 67, 63–69. of slag: a review. Appl. Geochem. 57, 236–266.
Basavarajappa, M., Miskovic, S., 2016. Investigation of gas dispersion characteristics in Potgieter, J.H., Kabemba, M.A., Teodorovic, A., Potgieter-Vermaak, S.S., Augustyn, W.G.,
stirred tank and flotation cell using a corrected CFD-PBM quadrature-based moment 2006. An investigation into the feasibility of recovering valuable metals from solid
method approach. Miner. Eng. 95, 161–184. oxide compounds by gas phase extraction in a fluidised bed. Miner. Eng. 19,
Benyahia, S., 2008. Verification and validation study of some polydisperse kinetic the- 140–146.
ories. Chem. Eng. Sci. 63, 5672–5680. Potysz, A., van Hullebusch, E.D., Kierczak, J., 2018. Perspectives regarding the use of
Carranza, F., Romero, R., Mazuelos, A., Iglesias, N., Forcat, O., 2009. Biorecovery of metallurgical slags as secondary metal resources – a review of bioleaching ap-
copper from converter slags: Slags characterization and exploratory ferric leaching proaches. J. Environ. Manage. 219, 138–152.
tests. Hydrometallurgy 97, 39–45. Romeo, L.M., Díez, L.I., Guedea, I., 2011. Design and operation assessment of an oxyfuel
Chao, Z., Wang, Y., Jakobsen, J.P., Fernandino, M., Jakobsen, H.A., 2011. Derivation and fluidized bed combustor. Exp. Therm Fluid Sci. 35, 477–484.
validation of a binary multi-fluid Eulerian model for fluidized beds. Chem. Eng. Sci. Rong, L.W., Dong, K.J., Yu, A.B., 2014. Lattice-Boltzmann simulation of fluid flow
66, 3605–3616. through packed beds of spheres: Effect of particle size distribution. Chem. Eng. Sci.
Chen, X., Wang, J., 2014. A comparison of two-fluid model, dense discrete particle model 116, 508–523.
and CFD-DEM method for modeling impinging gas–solid flows. Powder Technol. 254, Sarfo, P., Das, A., Wyss, G., Young, C., 2017. Recovery of metal values from copper slag
94–102. and reuse of residual secondary slag. Waste Manage. 70, 272–281.
Cheng, Y., Zhang, W., Guan, G., Fushimi, C., Tsutsumi, A., Wang, C.-H., 2014. Numerical Shen, H., Forssberg, E., 2003. An overview of recovery of metals from slags. Waste
studies of solid-solid mixing behaviors in a downer reactor for coal pyrolysis. Powder Manage. 23, 933–949.
Technol. 253, 722–732. Singh, A., Verma, R., Kishore, K., Verma, N., 2008. Multi-stage fluidized bed column:
Coroneo, M., Mazzei, L., Lettieri, P., Paglianti, A., Montante, G., 2011. CFD prediction of hydrodynamic study. Chem. Eng. Process. 47, 957–970.
segregating fluidized bidisperse mixtures of particles differing in size and density in Stroh, A., Alobaid, F., von Bohnstein, M., Ströhle, J., Epple, B., 2018. Numerical CFD
gas–solid fluidized beds. Chem. Eng. Sci. 66, 2317–2327. simulation of 1 MWth circulating fluidized bed using the coarse grain discrete ele-
Das, M., Meikap, B.C., Saha, R.K., 2008. Characteristics of axial and radial segregation of ment method with homogenous drag models and particle size distribution. Fuel
single and mixed particle system based on terminal settling velocity in the riser of a Process. Technol. 169, 84–93.
circulating fluidized bed. Chem. Eng. J. 145, 32–43. Sun, Q., Lu, H., Liu, W., He, Y., Yang, L., Gidaspow, D., 2005. Simulation and experiment
Gyurov, S., Kostova, Y., Klitcheva, G., Ilinkina, A., 2011. Thermal decomposition of of segregating/mixing of rice husk-sand mixture in a bubbling fluidized bed. Fuel 84,
pyrometallurgical copper slag by oxidation in synthetic air. Waste Manage. Res. 29, 1739–1748.
157–164. Sun, X., Waters, K.E., 2014. Development of industrial extractants into functional ionic
Hassanzadeh, A., Firouzi, M., Albijanic, B., Celik, M.S., 2018. A review on determination liquids for environmentally friendly rare earth separation. ACS Sustain. Chem. Eng. 2,
of particle–bubble encounter using analytical, experimental and numerical methods. 1910–1917.
Miner. Eng. 122, 296–311. Syamlal, M., 1987. The particle-particle drag term in a multiparticle model of fluidiza-
Igci, Y., Sundaresan, S., 2011. Constitutive models for filtered two-fluid models of flui- tion, Topical Report, National Technological Information Service, DOE/MC/21353-
dized gas–particle flows. Indus. Eng. Chem. Res. 50, 13190–13201. 2373 (DE87 006500).
Jalali, P., Hyppänen, T., 2010. Verification of continuum models for solids momentum Van Dyk, L., Mariba, E.R.M., Chen, Y., Potgieter, J.H., 2010. Gas phase extraction of iron
transfer by means of discrete element method. Indus. Eng. Chem. Res. 49, from its oxide in a fluidized bed reactor. Miner. Eng. 23, 58–60.
5270–5278. Wang, G., Ge, L., Mitra, S., Evans, G.M., Joshi, J.B., Chen, S., 2018. A review of CFD
Jia, W.C., Hays, R., Findlay, J.G., Knowlton, T.M., Karri, S.B.R., Cocco, R.A., Hrenya, modelling studies on the flotation process. Miner. Eng. 127, 153–177.
C.M., 2012. Cluster characteristics of Geldart group B particles in a pilot-scale CFB Wang, J., van der Hoef, M.A., Kuipers, J.A.M., 2013. Comparison of two-fluid and discrete
riser II. Polydisperse systems. Chem. Eng. Sci. 68, 72–81. particle modeling of dense gas-particle flows in gas-fluidized beds. Chem. Ing. Tech.
Karimi, M., Akdogan, G., Bradshaw, S.M., 2014a. A CFD-kinetic model for the flotation 85, 290–298.
rate constant, Part II: model validation. Miner. Eng. 69, 205–213. Ye, G., Burström, E., Kuhn, M., Piret, J., 2003. Reduction of steel-making slag for recovery
Karimi, M., Akdogan, G., Bradshaw, S.M., 2014b. A computational fluid dynamics model of valuable metals and oxide materials. Scand. J. Metall. 32, 7–14.
for the flotation rate constant, part I: model development. Miner. Eng. 69, 214–222. Zhang, L., Zhang, L.N., Wang, M.Y., Li, G.Q., Sui, Z.T., 2007. Recovery of titanium
Le Lee, J., Lim, E.W.C., 2017. Comparisons of eulerian-eulerian and CFD-DEM simulations compounds from molten Ti-bearing blast furnace slag under the dynamic oxidation
of mixing behaviors in bubbling fluidized beds. Powder Technol. 318, 193–205. condition. Miner. Eng. 20, 684–693.
Liu, J., Yu, Q., Zuo, Z., Duan, W., Han, Z., Qin, Q., Yang, F., 2016. Experimental in- Zhang, Y., Zhao, Y., Lu, L., Ge, W., Wang, J., Duan, C., 2017. Assessment of polydisperse
vestigation on molten slag granulation for waste heat recovery from various me- drag models for the size segregation in a bubbling fluidized bed using discrete par-
tallurgical slags. Appl. Therm. Eng. 103, 1112–1118. ticle method. Chem. Eng. Sci. 160, 106–112.
Lu, H., Gidaspow, D., 2003. Hydrodynamics of binary fluidization in a riser: CFD simu- Zhou, Q., Wang, J., 2015. CFD study of mixing and segregation in CFB risers: Extension of
lation using two granular temperatures. Chem. Eng. Sci. 58, 3777–3792. EMMS drag model to binary gas–solid flow. Chem. Eng. Sci. 122, 637–651.
Lu, H., Gidaspow, D., Manger, E., 2001. Kinetic theory of fluidized binary granular
mixtures. Phys. Rev. E 64, 061301.

You might also like