Accepted Manuscript Fluidized beds modeling: Validation of 2D and 3D simulations against experiments
Hongbo Shi, Alexandra Komrakova, Petr Nikrityuk PII: DOI: Reference:
S0032-5910(18)30947-1 https://doi.org/10.1016/j.powtec.2018.11.043 PTEC 13878
To appear in:
Powder Technology
Received date: Revised date: Accepted date:
13 December 2017 20 October 2018 10 November 2018
Please cite this article as: Hongbo Shi, Alexandra Komrakova, Petr Nikrityuk , Fluidized beds modeling: Validation of 2D and 3D simulations against experiments. Ptec (2018), https://doi.org/10.1016/j.powtec.2018.11.043
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT Fluidized Beds Modeling: Validation of 2D and 3D Simulations Against Experiments Hongbo Shi1, Alexandra Komrakova1*, Petr Nikrityuk2 1
Department of Mechanical Engineering, University of Alberta, Edmonton, Alberta, Canada
T
2
IP
Department of Chemical and Materials Engineering, University of Alberta,
Abstract
CR
Edmonton, Alberta, Canada
US
The flow field in fluidized beds is studied numerically using a two-phase Eulerian-Eulerian model. The main goal of the study is to demonstrate the importance of the numerical model
AN
setup: we investigate the effect of model dimensionality (2D & 3D), the impact of the flow regimes (laminar & turbulent), the choice of the model parameters (specularity coefficient
M
responsible for particle-wall interaction and restitution coefficient characterizing particle-particle interaction), mesh resolution, and gas-solid drag sub-models
ED
(Syamlal-O’Brien and Gidaspow). The results of simulations are validated with the experimental data obtained by Taghipour et. al. (Chem. Eng. Sci., 60(24), 2005,
PT
6857-6867). The comparison of the results shows that the 2D simulations tend to overpredict the experimental data in terms of void fraction distribution across the flow.
CE
When the 3D simulations are used, the numerically obtained profiles of the void fraction inside the bed agree well with experimentally measured values. The results of the 3D
AC
simulations clearly illustrate that the finite size of the domain where fluidization occurs plays a significant role in the size of bubbles and in the spatial distribution of the volume fraction of particles inside fluidization zone. It is shown that turbulence has a minor influence on the fluidized bed structure and void fraction distribution. However, it significantly affects the gas phase velocity in the upstream region of a reactor. A thorough study of numerous numerical model setups concludes that for the 2D simulations of fluidized beds there is a combination of model parameters such as specularity and
*
Corresponding author. E-mail:
[email protected]
ACCEPTED MANUSCRIPT restitution coefficients, numerical grid resolution, and discretization scheme for convective terms that will always produce a good agreement between numerical and experimental results. Only the 3D simulations allow obtaining physically realistic results that are less sensitive to a choice of numerical parameters.
Keywords: Fluidized beds, validation, 2D, 3D, turbulence, restitution coefficient,
IP
T
specularity coefficient
CR
Nomenclature
US
drag function
CD
d s , d p diameter of solid particle
activation energy
ess
coefficient of restitution
g
gravitational acceleration
g 0,ss
radial distribution function
K
specularity coefficient
kg
turbulence quantities of gas phase
ks
turbulence quantities of solid phase
K pq
interphase momentum exchange coefficient
K gs
interphase momentum exchange coefficient between gas and solid phase
CE
PT
ED
M
AN
EA
nt
AC
k K-Epsilon turbulence model temperature exponent
p
pressure shared by all phases
ps
solid pressure
R pq
interaction force between phases
Res
Reynolds number for solid particles
RNG renormalization group method
ACCEPTED MANUSCRIPT superficial gas velocity
Ug
volume averaged velocity for gas phase
Vq
volume of phase q
Vsb
volume of the spouted bed reactor
vq
velocity of phase q
vg
velocity of gas phase
vs
velocity of solid phase
vq
particle velocity fluctuation
US
CR
IP
T
U
Greek letters
q
volume fraction of q phase
AN
s ,max maximum packing limit
volume fraction of solid phase
g
volume fraction of gas phase
q
physical density of phase q
g
physical density of phase g
ˆ q
effective density of phase q
q
q th phase stress-strain tensor
q
bulk viscosity of phase q
s
gs
ED
PT
CE
AC
M
s
collisional energy dissipation kinetic energy transfer
s
granular temperature
fluid-particle interaction coefficient
s
solid shear viscosity
s ,col
solid collisional viscosity
s ,kin solid kinetic viscosity
frictional viscosity
g
viscosity of gas phase
s
volume fraction of solid phase
q
granular temperature
t , g
turbulent viscosity
g
turbulence dissipation rate of gas phase
IP
s , fr
T
ACCEPTED MANUSCRIPT
k ,g , z,g
influence of dispersed phases on continuous phase
standard error of the estimate
US
est
CR
k , z adjustable constants
p
particle
s
solid phase
M
gas phase
ED
g
Introduction
PT
1
AN
Subscripts
Fluidized beds is a widely used technology in high-temperature conversion of materials including
CE
gasification, waste treatment, and CO2 utilization [1]. Analysis of different types of fluidized beds and their application in industry shows that the next challenging task for engineers is to
AC
increase the efficiency of fluidized beds. This can be done by means of computational fluid dynamics (CFD) modeling of processes occurring inside the fluidized beds. Recently, computer modeling became one of the most crucial elements in the design and optimization of novel technologies in the field of industrial engineering. One of the advantages of computer mode ling is that the behavior and characteristics of the fluidized bed reactor can be investigated without fabrication of a pilot-scale experimental prototype. Therefore, the total costs of product development and design optimization are reduced significantly. However, the numerical simulations of multiphase flows in fluidized beds often involve complex physical and chemical phenomena which have to be modeled using the advanced mathematical models implemented into
ACCEPTED MANUSCRIPT a CFD software. To guarantee adequacy and reliability of any model or software, verification and validation processes are required [2, 3, 4]. The verified and validated numerical models can be used to study the effects of different parameters on the efficiency of the process when experiments are challenging and expensive. With an impressive development in commercial CFD software, e.g., ANSYS FLUENT [5], ANSYS CFX [6], STAR-CCM+ [7], FLOW-3D [8], and in multi-phase flow models, it
T
became possible to simulate the three-dimensional (3D) fluidized beds without oversimplification.
IP
The Eulerian-Eulerian (E-E) and Eulerian- Lagrangian (E-L) are the two main numerical
CR
approaches used in CFD-based fluidized bed models. In the E-E approach, the fluid and solid phases are considered as interpenetrating continuum phases [9]. The conservation equations are
US
required to describe the transportation of mass, momentum, and energy. The E-L approach treats the fluid phase as continuum and the solid as individual dispersed particles. The Newton’s law of
AN
motion and Navier-Stokes equations are used to solve the discrete phase and the fluid phase, respectively [9].
M
Before we proceed with the literature review of the E-E simulations of fluidized beds, we will outline and define the most important parameters (except properties of particles, e.g. density,
ED
size and shape) that determine the hydrodynamics of the gas-solid flow in a fluidized beds, and, therefore, should be accounted for in the numerical model. These parameters are the superficial
PT
gas velocity, the gas-solid drag, the specularity and restitution coefficients, and the frictional viscosity.
CE
The superficial gas velocity represents the inlet velocity of gas through a packed bed in the gasifier. The fluid regime can be changed with the increase of the upward superficial gas velocity
AC
from zero to the point of fluidization. In this case, the upward gas-solid drag force exceeds the downward body force of gravity. The gas-solid drag force (or just the drag) is the resistance force caused by the motion of a particle through a fluid [5]. The drag is the main reason for the bed expansion and intensified mixing between particles and gas. Due to the complex interactions between particle-particle and particle-wall, the description of dynamic behavior of the fluidized bed is complicated. The specularity coefficient is used to characterize particle-wall interaction. The restitution coefficient is used to characterize the particle-particle interaction. Both of them play an important role in the bed behavior. Different hydrodynamic behavior (e.g. flow structure, pressure drop, or particle velocity) can be captured by the model using different values of
ACCEPTED MANUSCRIPT specularity coefficient and restitution coefficient [10, 11]. After collision between the particles, the stress is generated due to the enduring contact and momentum transfer through friction between the particles which is called as frictional viscosity [12]. The CFD-based models to simulate fluidized beds can reveal the influence of each of these parameters and guide the improvement of the process. Even though there are numerous CFD studies exploring the influence of the restitution
IP
T
coefficient ( ess ) and the specularity coefficient ( K ) on the flow behaviour inside the fluidized beds, e.g. see [10, 11], a clear graphical explanation of these coefficients is still needed. For that
CR
reason, we introduce Fig. 1 (a), which shows the definition of the restitution coefficient (particle-particle collision) in the fluidized bed. For perfectly inelastic collision, ess 0 , the
US
particle A and the particle B stick and move together with a final speed in the bed after collision. For perfectly elastic collision, ess 1 , the particle A and the particle B bounce away from each
AN
other and move separately with the same magnitude of final speed after collision. The restitution coefficient ess is defined as follows [13]:
M
v1 v2 u1 u2
(1)
ED
ess =
where v1 and v2 are the post-collision velocities of particle A and particle B, respectively. The
PT
u1 and u2 are the pre-collision velocities of particle A and particle B, respectively. Fig. 1 (b) illustrates the meaning of the specularity coefficient that defines particle-wall
CE
collision in the fluidized bed. For perfectly specular collision, K 0 , the particle collides with the wall and then bounces to the central region of bed with the same incidence angle and the reflection
AC
angle ( 1 = 2 ). For perfectly diffuse collusion, K 1 , the particle has different incidence and reflection angles ( 1 2 ) after collision with the wall. With these definitions in mind, we reviewed published studies on numerical simulations of fluidized beds. Many researchers studied the characteristics of gas-solid fluidized bed reactors by applying the E-E and E- L models in the CFD software, ANSYS FLUENT. Taghipour [14] used the E-E approach to study the effect of different drag sub-models (Syamlal-O’Brien [15], Gidaspow [16], and Wen-Yu [17]) when running the two-dimensional (2D) simulations. The comparison of the numerical and experimental bed pressure drop, bed expansion ratio, and
ACCEPTED MANUSCRIPT gas-flow pattern revealed a good agreement between the numerical and experimental results. Loha et al. [10] carried out an E-E simulation to investigate the effect of specularity coefficient, K , on the hydrodynamic behavior of the bed for 2D simulations. Different values of restitution coefficient ess = (0, 0.01, 0.1, 0.3, 0.6 and 1.0) were tested. It was shown that with the decrease of K from 1.0 to 0.01 , the pressure drop across the bed is constant and remains in
T
the range 1.90 1.92 kPa. However, the bed pressure drop reduces significantly to 1.84 kPa
IP
when K = 0 . The comparison of the numerical and experimental data of time-averaged axial particle velocity and granular temperature showed a good agreement if specularity coefficient is
CR
equal to 0.1 , 0.3 , 0.6 and 1.0 .
Loha [11] used the E-E approach to study the influence of the restitution coefficient, ess ,
US
on the hydrodynamics of the fluidized beds using 2D laminar model. The effect of the restitution coefficient ess in the range (0.85 1.0) on the time-averaged solid volume fraction is not clear.
AN
The assumption of ess 1 results in the flow pattern without any bubbles in the bed which is
M
physically non-realistic. When ess decreases from 1.0 to 0.85 , the quantity and the size of bubbles increases and the averaged solid volume fraction decreases accordingly. The large
ED
fluctuation in the pressure drop within the bed is contributed to the formation of a large number of big bubbles. The numerical results of time-averaged axial particle velocity and granular
PT
temperature were in line with the experimental measurements for ess = 0.95 and 0.99 . Both E-E and E-L models were used by Almohammed [18] to study the effects of the flow
CE
rate of the injected gas on the 3D dispersed turbulence model. The numerical result showed that the restitution coefficient of 0.6 and specularity coefficient of 0.5 with Syamlal-O’Brien
AC
drag function give the best agreement with the experimental data on the bed height and the equivalent bubble diameter over time. The E-E approach can successfully predict the flow pattern at a mass flow rate of 0.005 kg/s, while a significant deviation is produced as the flow rate increases to 0.006 kg/s. Stroh [9] studied and compared three different numerical methods, Euler-Euler (E-E), Euler-Lagrange-stochastic (E-L-S), and Euler- Lagrange-deterministic (E-L-D) on a 3D laminar model. The comparison of numerical and experimental measured bed expansion and equivalent bubble diameter shows that all three approaches can predict the bubble development for the inlet air mass flow rates of 0.005 and 0.006 kg/s. For the E-E approach, the accurate flow pattern can be
ACCEPTED MANUSCRIPT predicted for the smaller mass flow rate (0.005 kg/s) but not for the higher mass flow rate (0.006 kg/s) between the time interval of 0 300 ms. Cornelissen [19] used the E-E approach to investigate different mesh size, time step, and convergence criteria on a 2D laminar model. The comparison of numerical and experimental data on overall bed voidage shows that the model using 60 500 grid gives a small deviation of 4% in comparison of 45 250 , 50 400 , and 70 700 grids. When time step t = 0.001 s is used, the
IP
T
overall bed voidage values predicted numerically agree well with the experimental values. Jin [20] used the E-E approach to investigate the effect of particle-particle restitution
CR
coefficient, the particle-wall restitution coefficient, and the specularity coefficient on a 3D k dispersed model. The large particle size ( d p 100 m ) leads to a high sensitivity of
US
particle-particle restitution coefficient on gas and solids velocity and solid volume fraction distribution in comparison to a small particle size ( d p 100 m ). The sensitivity of
AN
particle-particle restitution coefficient on the variable particle density was not explained. The particle-wall restitution coefficient has a negligible effect on the time-averaged solid volume
M
fraction, axial particle and gas velocity. The solids volume fraction decreases from 14% to 4% near
ED
the wall when the specularity coefficient was increased from 0 to 0.01. The comparison between numerical and experimental axial pressure gradient shows that the particle-particle restitution coefficient of 0.99, particle-wall restitution coefficient of 0.99, and specularity coefficient of 0 can
PT
give the best result with an averaged deviation from the experiment of
5.6% .
In series of works Ghoniem and co-workers [21, 22] carried out 2D and 3D numerical
CE
simulations of dense solid–gas cylindrical and rectangular fluidized beds using the E-E based model utilizing the MFIX code. The influence of different wall boundary conditions and drag
AC
models and the specularity coefficient were extensively studied numerically. In particular, it was shown that in the case of thin rectangular fluidized beds the effects of the front and the back walls are significant [22]. The influence of the specularity coefficient, which is responsible for the momentum transfer from the particles to the wall, showed that the appropriate value of the specularity coefficient depends strongly on the superficial gas velocity of the bed. However, it should be noted that no turbulence models were used in that simulation [22]. Li and co-workers [23] investigated numerically the influence of the solid-phase wall boundary conditions on the hydrodynamics of a bubbling fluidized bed. Parametric studies of the
ACCEPTED MANUSCRIPT particle–wall restitution coefficient and specularity coefficient were performed to evaluate their impact on the predicted fluidized bed behavior in terms of bed expansion, local voidage, and solid velocity. Both 2D and 3D simulations (using the MFIX code) were conducted and compared with available experimental data. Good agreement was observed. However, it should be noted that no any turbulence models were used in that simulation. Kong et al. [24] evaluated numerically the effect of wall boundary conditions on the
T
hydrodynamics of circulating fluidized beds. The Eulerian-Eulerian 2D CFD-based model
IP
(ANSYS Fluent 13.0) coupled with k turbulence model formulated for each phase was used.
CR
The results showed that the specularity coefficient has a strong effect on the solid particle distribution near the wall. At the same time, the particle–wall restitution coefficient has less
US
effect on solid particle distribution near the wall than that of the specularity coefficient. Numerous amount of studies on the Eulerian-Eulerian models applied for the fluidized bed
AN
simulations show that these models are extensively used. The major reason for this is that they are less computationally demanding as compared to other more sophisticated modeling approaches
M
where each particle motion is being resolved. The potential applicability of the E-E approach is high since the unlimited number of particles can be modeled with this approach. The disadva ntage
ED
of the E-E approach is that some discrete character of the solid phase cannot be identified [25]. The additional source terms (closure laws) are required in the E-E approach to define the interactions
PT
between the solid and fluid phases. The advantage of the E-L approach is that the high- level details of the solid phase are available since the individual particles are tracked. On the other hand,
CE
significant computational cost (e.g. CPUs and memory resources) is required for the E-L approach to give a statistically meaningful numerical prediction of the real hydrodynamic behavior [25].
AC
The number of particles that can be modeled using the E-L approach is limited in the range from 106 to 109 [26]. The number of particles in each grid cell is also an important parameter. Due to
the validity of the drag force closure relations used in the E-E approach, the particle size must be smaller than the grid cell size: the edge of the cell in each direction should be at least two particle diameters ( x 2d p , y 2d p , and z 2d p ). Table 1 shows the number of particles in each grid cell for several works. As can be seen, several studies do not follow the requirement o f the particle resolution. These studies are highlighted in the table. The numerical results from these works might not replicate the physical behavior. The Syamlal-O’Brien and Gidaspow drag sub- models cannot be applied in this cases due to less number of particles ( < 10 ) in each cell [26].
ACCEPTED MANUSCRIPT In spite of the numerous works on validation of the CFD-based models (E-E or E-L) applied to simulations of fluidized bed, there is a lack of comparison of the results obtained on 2D and 3D simulations specifically for a thin rectangular bed. Often in the literature, the 2D simulations are used to extrapolate theoretical findings for different types of fluidized beds, thin to wide [27]. However, in the case of a thin rectangular fluidized bed, the wa ll effects in the third dimension might be significant. The purpose of this work is to show the significance of the third
T
dimension on the simulation results. The experimental work of Taghipour et al. [14] is used as a
IP
reference. Our goal is to estimate the magnitude of the difference between 2D and 3D simulations
CR
and compare the numerical and experimental results. A similar effect was investigated by Ghoniem and co-workers [22] but only for the laminar flows. We will consider turbulent models as
US
well. Xie et. al. [28, 29] studied numerically two- and three-dimensional computational modeling of the fluidized beds. A good agreement between 2D and 3D simulations was observed for
AN
bubbling regimes with low gas velocity. For slugging and turbulent regimes, the 2D and 3D simulations produced different results. However, we note that the Authors did not carry out any
M
validation against experimental data.
Motivated by a lack of solid comparison of 2D & 3D simulation results to experimental
ED
data, the main objective of the present study is to validate the 2D & 3D Euler- Euler-based models available in a commercial CFD software, against experimental data [14]. It is also necessary to
PT
analyze the effects of the laminar and turbulent k flow models on numerical results. Together with the influence of a grid resolution, the sensitivity of key model parameters (the specularity
CE
coefficient, the restitution coefficient, and the drag sub-model) will be evaluated. The rest of the paper is organized as follows. Section 2 presents the description of the
AC
problem. The details of reference experiments are given in this section. The computational model is described in Section 3. Section 4 outlines the results of the present study. Finally, the conclusions are drawn in Section 6.
2
Problem description
To evaluate the capabilities of a commercial CFD software to capture the physics of fluidized beds and study the sensitivity of parameters of a fluidized bed, the experiment of Taghipour [14] is numerically replicated. The principal scheme of the experiment and simulation is shown in Fig. 2.
ACCEPTED MANUSCRIPT According to the experimental setup, the Plexiglas column has a height of 1.5 m, a width of 0.28 m, and a thickness of 0.025 m. Spherical glass particles with a diameter of 250 300 m , density
2500 kg / m3 are fluidized with air at ambient conditions. The static bed height is 0.4 m with a solid volume fraction of 0.6. The maximum packing limit for the solid phase is max 0.63 . In the experiment, the overall pressure drop and bed expansion were monitored at a superficial gas
T
velocity of U = 0.38 m/s and 0.46 m/s. The measurements of time average pressure drop were
IP
recorded at 10 Hz for 20 s intervals once the steady-state conditions were achieved. To compare
CR
the numerical results with the experiment, the voidage measurements were conducted at the height
3
US
0.20 m above the distributor plate across the width of the column.
omputational model
AN
On the basis of the Euler-Euler approach, the unsteady laminar and turbulent Reynolds-averaged Navier-Stokes (RANS) formulations for the mass and momentum are solved for gas and solid
M
phase separately. The gas phase is modelled using the standard k - turbulence model with a standard wall function. The solid phase is modeled with the kinetic theory of granular flows
ED
(KTGF) [5] that describes the fluctuations and collisions between the particles. The momentum equation for the gas and solid phases is given by the modified Navier-Stokes equations that include
PT
the inter-phase momentum transfer terms. In particular, the frictional viscosity, the drag coefficient sub- model, the granular temperature, the bulk viscosities, and the granular temperature
CE
are the additional terms that depict the momentum exchange. Both turbulence interaction and turbulent dispersion between the gas and solid phases are estimated by the Simonin & Viollet
AC
theory [30]. The Tchen theory [31] of dispersion of discrete particles is utilized in the dispersed turbulence model to estimate turbulent quantities (dispersion coefficients, correlation functions, and the turbulent kinetic energy) for the particle phase. The Syamlal-O’Brien [15] and Gidaspow [16] drag sub- models are used. The Syamlal-O’Brien drag sub- model is based on the terminal particle velocities in fluidized beds [32]. The Gidaspow drag model is an integration of the Wen-Yu model and the Ergun model that can be applied for the flow where viscous forces are dominant [33]. As the inlet boundary condition, a superficial gas velocity of 0.38 m/s or 0.46 m/s is set. The outlet boundary condition is a fully-developed flow. The boundary conditions for turbulence
ACCEPTED MANUSCRIPT are specified with uniform value, 5% for turbulence intensity and 0.1 for turbulence viscosity ratio. The boundary conditions in terms of specularity coefficients (0.01, 0.1, 0.5, and 0.9) and the particle-particle restitution coefficients (0.90, 0.95, and 0.99) are applied in the simulation and their effects are studied in detail. The simulations were performed using the CFD commercial software, ANSYS FLUENT 16.2. The computational geometry was discretized with three block-structured grids using a finite
T
volume method with 56 300 , 112 600 , and 56 300 5 cells (see Table 2). The particle
IP
diameter considered in simulations was d p = 275 m .
CR
The pressured-based coupled scheme [5] was used to solve mass and momentum conservation equations corresponding to the behavior of the fluidized bed. To fulfill the coupling
US
algorithms, the implicit discretization of both pressure gradient terms and face mass flux were included in the equation. Quadratic upwind interpolation for convection kinematics (QUICK)
AN
scheme [34] was activated to discretize the convection terms in the momentum equations. The modified type of the high-resolution interface capturing (HRIC) scheme [35] was used to estimate
M
the volume fraction of gas or solid. In comparison between the QUICK and the first-order upwind schemes, the accuracy of numerical solution can be improved significantly by the modified HRIC
ED
scheme since the non- linear blend of upwind and downwind differencing are included in this scheme [35]. The second-order upwind scheme for the discretization of the convective terms was
PT
used in the k - model. The details of the models and schemes used in the numerical simulation of fluidized beds are given in Table 3.
CE
The total physical simulation time of 20 seconds was completed for each simulation. The time-averaged values were obtained from the last 18 seconds after a dynamic steady state
AC
converged solution was reached within two seconds after the process initiation. To avoid instability, a small constant time step of 104 s was used for each case, and the number of inner iterations on each time step was set to 50. The total of 31 simulations (eight 3D cases and twenty three 2D cases) were carried out in this study. The cases are described in Table 4. Finally, it should be noted that in all 2D and 3D simulations we used the full bed according to its sizes in 2D and 3D cases.
4
Results
ACCEPTED MANUSCRIPT To gain a general understanding of the accuracy of the numerical predictions of voidage fraction, the standard deviation profiles of the 2D and 3D simulations were estimated and are shown in Fig. 3. The formula of standard deviation [36] is shown below:
est % =
(Y
simulation
Yexperiment )2
N
100
(2)
T
where est is the standard error of the estimate, Yexperiment and Ysimulation are the experimental and
IP
predicted points of the voidage fraction along x coordinate at z = 0.2 m, respectively. The standard deviation provides reliable measure of error as compared to other functions (e.g. weight
CR
standard error) since this method comes from Linear Least Squares Regression used by curve fitting. The standard deviation of 3D and 2D models is in the range [3.85 8.54] % and
US
[3.97 13.08] %, respectively. The difference between the maximum and minimum deviations for
3D and 2D simulations is 4.69% and 9.10%. This reveals that the results of the 3D simulations of
AN
fluidized bed agree better with the experimental results, and therefore, should be used. However, a proper choice of parameters in 2D simulations might also lead to results that match experiments.
M
The detailed analyses of the results is presented in the following order. First, we assess the
ED
difference between two- and three-dimensional simulations. Second, the results for the laminar and turbulent models are discussed. Effects of the coefficient of restitution and specularity
4.1
CE
viscosity sub-models.
PT
coefficients are considered next. Finally, we examine the effects of the drag force and the frictional
Three-dimensional versus two-dimensional approaches
AC
This section compares the numerical predictions of two different modelling approaches, 3D and 2D. For a 3D model, the total simulation time of 2 weeks is required to simulate the physical process time of 20 seconds. However, a 2D model needs only 2
3 days to simulate the same
physical time. To illustrate the fluidized bed behavior, Fig. 4 shows the snapshots at different time instances of the slice of a 3D contour plot of the solid volume fraction and the turbulent viscosity ratio. The red color represents the maximum solid volume fraction and turbulent viscosity ratio, whereas the blue color denotes the minimum value of zero. As can be seen from Fig. 4 (a), bubbles formed at the bottom of the bed move up to the bed surface with the increasing bubble size. The
ACCEPTED MANUSCRIPT shape of the bubbles ceases to remain spherical because of the intensive breakage and coalescence. Upon reaching the bed surface, the solid particles are forced towards the wall with the burst of big bubbles. Then the particles fall down along the walls due to the gravitational force. This behaviour of particles can be seen in Fig. 5 that presents a time-averaged solid volume fraction in the middle slice of the domain. A lower solid volume fraction (30%) is observed in the central area of the
T
bed while the higher solid volume fraction of 60% is achieved in the regions close to the wall. The time snapshots of the turbulent viscosity ratio for the same 3D simulation case are
IP
depicted in Fig. 4 (b). The turbulent viscosity ratio – defined as a ratio between turbulent t and
CR
molecular o dynamic viscosities – is necessary to estimate the level of turbulence within the simulation domain. As can be seen from the Fig. 4, the distribution of turbulence is consistent with
US
the gas flow patterns. High turbulent viscosity ratio (t / o > 10) exists inside the big bubble space at the startup phase ( t < 2.0 s). After reaching the dynamic steady state ( t > 2.0 s), weak
AN
turbulent viscosity ratio in the range (0 10) is observed inside the bed, while strong turbulence
M
in the range (15 110) is distributed above the bed surface. This indicates that only a small amount of gas (bubbles) is trapped inside the bed. The gas primarily fills the space above the bed
ED
surface by releasing the trapped gas from the bubbles. To study the transient dynamics of the gas and solid flows in a fluidized bed, we use the
flow velocity U s [37]:
PT
volume-averaged or global velocity of the solid phase in the entire bed. The volume-averaged solid
1 V fb
CE
Us =
x
y
0
0
H
0
us2, x us2, y us2,z d x d y d z
(3)
AC
where V fb is the volume of the fluidized bed reactor and us is the velocity of the solid phase. This global velocity is used to calculate a spinup or startup time for the volume- force driven flows [37]. Time histories of U s calculated using different values of specularity coefficient for 2D and 3D RANS approaches are depicted in Fig. 6. A noticeable narrow peak in the volume-averaged velocity profile within the first 2 seconds is captured in all cases. This peak corresponds to the startup time. Therefore, the fluidized bed reaches a dynamic steady state within two seconds of process initiation. After the startup period, steady fluctuations of solid volume fraction are observed in the bed surface that might be explained by continuous formation, rise, coalescence, and the breakup of bubbles.
ACCEPTED MANUSCRIPT Analysis of U s time history shown in Fig. 6 reveals that the solid velocities predicted using the 2D model are higher than velocity obtained from the 3D model. Dominant velocity in the ranges 0.2 0.3 m/s and 0.1 0.2 m/s are found for 2D and 3D cases, respectively. The lower solid velocity from the 3D models is explained by the loss of extra friction due to the increased contact area between the fluid and the walls. To capture this phenomenon, the simulation of the fluidized bed has to use the 3D approach rather than the 2D approach to obtain physically realistic
IP
T
results. Even if the 2D grid is refined from 56 300 to 112 600 , the numerical results do not approach the experimental measurements. However, significant computational demands
CR
associated with the 3D simulations make 2D simulations more appealing for the sensitivity analyses of the parameters. For that reason, it is necessary to assess to what extent the 2D results
US
can be used for these purposes.
Since the main goal of this study is to validate the numerical approach, we will start
AN
analyses of the 3D numerical results that replicate the real phenomena better than 2D simulations. Fig. 7(a) compares the predicted void fraction values against the experiment data along x
M
coordinate at z = 0.2 m with the inlet gas velocity of 0.38 m/s. The experimental data is adapted from the experiment carried out by Taghipour [14]. The averaged deviation from the experimental
ED
values for all three numerical calculations in Fig. 7 is 5.36% suggesting that the 3D approach produces a good agreement with the experimentally measured void fraction profiles. It can be seen
PT
that there is a flat distribution with high gas volume fraction in the central region and sharp distribution with low gas volume fraction close to the wall. The axial variations in gas volume
CE
fraction are similar for the coefficient of restitution 0.9 and 0.95 without activation of the fictional viscosity sub- model. The deviations from experimental data are 4.86% and 4.71% for these cases,
AC
respectively. It confirms that there is no qualitative change in the nature of variation of the results by varying the coefficient of restitution within a range of [0.9 0.95] . However, a relatively large deviation of 6.50% against experimental data is presented by the increase of specularity coefficient, K , from 0.1 to 1. This overestimated prediction is caused by a perfectly diffuse collision between the particle and the walls when K = 1 . Fig. 7 (b) depicts the 3D simulated time-averaged void fraction profile with the inlet gas velocity of 0.46 m/s. By comparing the numerical results with the experiment points, all four cases give small deviation in the range between 3.85% to 6.37%. With the increase of the specularity coefficient from 0.1 to 1, the minimum value of void fraction (41.5%) can be observed on the wall, the main possible reason is
ACCEPTED MANUSCRIPT explained in the previous discussion. Finally, Fig. 7 (c) presents time-averaged void fraction profiles predicted using 3D laminar model and 3D k - disperse turbulence model. It can be seen that the turbulence model produced better agreement with experiments.
4.2
Laminar model versus RANS model
T
The three-dimensional simulations of fluidized bed using an unsteady RANS k model show
IP
that the flow within the bed is laminar which is indicated by a low turbulent viscosity ratio (< 10)
CR
. At the same time, stronger turbulence develops above the bed surface. Therefore, it is necessary to assess what flow model (laminar or turbulent) should be used. The 2D fluidized bed configuration was simulated on grids Grid1 56 300 and Grid2 112 600 . The particle
US
concentration distributions inside the fluidized bed for the laminar and RANS models are shown in Fig. 8 for Grid1 and in Fig. 9 for Grid2, respectively. Similar to the results of the 3D simulation, in
AN
2D simulations, the bubbles undergo the formation, rise, coalescence, and burst. By comparing the flow pattern in both laminar (see Fig. 8 (a) and 9 (a) and RANS (see Fig. 8 (b) and 9 (b)) models, it
M
is noticed that the solid flow patterns inside the bed are relatively symmetrical for both models at
ED
the startup state (t < 2 s) . After reaching the dynamic steady state, similar bubble size and expanded bed height can be observed for both laminar and RANS models. In order to study the overall behavior of the fluidized bed, the 2D time-averaged solid
PT
fraction contour plots and gas velocity vector plots for laminar and RANS models on different grids are depicted in Fig 10 and 11. The symmetry in the solid pattern and gas velocity is achieved
CE
in all cases. The solid-phase distribution is similar for both laminar and RANS models in both girds suggesting that fluidization hydrodynamics is not sensitive to the choice of the flow model. A
AC
higher solid volume fraction of 0.6 is achieved close to the wall and at the bottom of the bed (0.45). A lower solid volume fraction of 0.3 is observed in the center of the bed. This distribution of solid volume fraction reveals that more bubbles are formed at the center region rather than at the bottom. The 15% solid volume fraction is observed at the bed surface demonstrating the burst of bubbles when they approach the surface. The analyses o f the distribution of fluidizing gas velocity (Fig. 10 (b) and Fig. 10 (b)) reveal that the fluidizing gas velocity in the center of the bed is high and gradually decreases above the bed surface. When the fluidizing gas flows down to the bottom, it changes direction after encountering the inlet gas and the magnitude of the velocity vector
ACCEPTED MANUSCRIPT decreases. Consequently, the mixture of fluidizing gas and inlet gas forms a vortex flow of gas in both sides between the center and the wall. This vortex is an important feature of the fluidized beds due to its governing role in heat and mass transfer applications. In particular, the residence time of solid particles in this zone is significantly larger than in the bulk zone. Fig. 12 displays the contour plots of the turbulent viscosity ratio predicted at different time instances using the 2D geometry with 56 300 and 112 600 mesh resolution. The observations
T
are similar to the 3D results: weak turbulence distributed in the bed, and high turbulence exists
IP
above the bed surface. In the startup state, we barely can see turbulence in the entire domain. After
CR
reaching the dynamic steady state, the turbulent viscosity ratio in the packed bed is in the region between 0 and 52, while in the gas phase is in between 103 and 310 for the model. As in the 3D
US
cases, the possible explanation is that less gas is being present in the bed. To illustrate qualitative differences between the laminar model and RANS model, Fig. 13
AN
depicts the time-averaged velocity of the gas phase along the center line corresponding to
x = 0.15 m. Higher gas velocity ( 0.25 1.7 m/s) exits in the bed while lower gas velocity (
M
0 0.5 m/s) distributed in the gas phase for both laminar model and RANS model. This can be explained by the presence of particles. In the gas phase, we obtain lower gas velocity between 0
ED
m/s to 0.38 m/s in the 2D RANS model and relatively higher velocity between 0.15 m/s to 0.44 m/s in the 3D RANS model. The reason for this phenomenon is that higher turbulent viscosity ratio
PT
(112 450) exists in the 2D model and the lower turbulent viscosity ratio (37-110) is distributed
in the 3D model.
CE
Referring to the comparison between solid phase velocities calculated using the laminar and turbulent models, Fig. 14 (a), (b) shows the time histories of the volume-averaged velocity of
AC
the entire solid phase predicted with laminar and RANS models, respectively. It can be seen that both laminar and RANS models show a similar trend of the solid velocity in the 2D model (in the range 0.2 0.3 m/s) and 3D model (in the range 0.15 0.25 m/s), which reveals that the turbulence does not play a significant role on solid velocity. Based on our simulation results that compare the laminar and turbulent models applied for the fluidized bed simulations, we conclude that the laminar model can be used when the isothermal fluidized bed is considered. For gasification and combustion applications where chemical reactions take place, the RANS model should be used. In these cases, species transport is affected by intensive turbulence above the bed surface since diffusion is different as compared to the
ACCEPTED MANUSCRIPT laminar flow.
4.3
Effect of coefficient of restitution ess
Fig. 15 (a) presents the axial distributions of the time-averaged void fraction for different values of coefficient of restitution ( ess = 0.90 , 0.95 , and 0.99 ) obtained as result of the 2D simulations.
T
The core-annular structure of the flow is predicted when collisions are inelastic. The void fraction
IP
is larger in the central region and less in the near wall region. A higher void fraction with a value of
CR
46% is observed on the wall for ess = 0.90 . This can be explained by the formation of many larger bubbles at a lower value of coefficient of restitution. The deviation of numerical results against
US
experiment data are 6.17%, 7.60%, and 7.72% for ess = 0.90,0.95, and 0.99 , respectively. This
4.4
AN
suggests that the void fraction is not sensitive to the choice of ess in the range [0.90 0.99] .
Effect of specularity coefficient K
M
Four different values of the specularity coefficients ( K ) in the range [0.01 0.9] are considered
ED
in the 2D numerical models. The comparison of the experimental and numerical data is presented in Fig. 15 (b). The differences in the numerical results using various specularity coefficients are more distinct while studying the time-averaged void fraction distribution. With the decrease of the
PT
specularity coefficient from 0.9 to 0.1, the trend of the void fraction with the deviation in the range
CE
between 7.57% and 7.99% can be observed. This suggests that the influence of K in the range between 0.1 to 0.9 on the numerical prediction of void fraction pattern is negligible. The existence of a core-annular structure of the flow was predicted for K between 0.1 and 0.9. When the
AC
specularity coefficient decreases to 0.01, a different gas volume fraction pattern is observed with the downflow of gas near the wall and at the center with the deviation of 9.68%. Two maximum gas volume fraction values, around 0.64, can be found at x = 0.05 and 0.225 m. The lower value of the specularity coefficient means the resistance between the particle and the wall is small. Consequently, more concentrated particles are distributed close to the wall, resulting in two peaks of the void fraction located each side between the center and the wall. Fig. 16 reveals that the solid velocity is significantly affected by the variation of specularity coefficient in the range between 0.1 and 1.0. For the superficial gas velocity of 0.38 m/s, the solid
ACCEPTED MANUSCRIPT velocity pattern of K = 1 model is lower than that of K = 0.1 model. The dominant solid velocity is around 0.1 m/s and 0.15 m/s for each model. The unity value of specularity coefficient leads to zero fluid velocity on the wall resulting in such a lower value of the time-averaged solid velocity. The effect of superficial gas velocity is more pronounced for the solid velocity profile. With the increase of the superficial gas velocity from 0.38 m/s to 0.46 m/s, the higher solid velocity pattern with an averaged value of 0.2 m/s is observed. Particles are carried up in the
T
central region and fall down in the near wall region by the inlet gas. Thus, a positive relation
4.5
CR
IP
between the superficial gas velocity and the particle velocity results in this phenomenon.
Effect of drag force and frictional viscosity
US
Fig. 17 (a) presents a comparison of the predicted time-averaged gas volume fraction with the experimental results by different parameters using Grid1. A better numerical result with a
AN
deviation of 4.50% is contributed from the RANS model by taking into account the Symal-Obrien drag sub- model and Johnson-et-al frictional viscosity. Whereas the laminar model gives the largest
M
discrepancy, 9.30%, with the experimental measurements. The simulation using Syamlal-Obrien drag submodel gives 6.17% deviation, while that using Gidaspow sub- model shows a deviation of
ED
8.78%. This reveals that use of the Syamlal-Obrien drag sub-model to predict the gas flow pattern in this fluidized bed gives more accurate results. The RANS model with a Syamlal-Obrien drag
PT
sub- model presents a relatively fully-developed bubbly flow pattern while the laminar model shows an annular flow pattern from the profile. This can be explained by irregular fluid flow from
CE
the turbulent regime and uncrossed orderly fluid flow produced by the laminar regime. The velocity of the fluid is not constant at any point of fluid for the turbulent regime, but constant
AC
velocity can be represented by the laminar regime. Consequently, the distribution of the void fraction pattern is consistent with the fluid velocity profile. Fig. 17 (b) shows the numerical results based on Grid2. The profile depicts that the simulation with the frictional viscosity can provide better agreement with the experimental data. The deviation for the frictional viscosity sub- model is 6.19%, while the deviation for the RANS model without applying frictional viscosity is 10.15%. A similar trend of the time-averaged void fraction profile can be predicted by different models without taking into account the frictional viscosity. The RANS model with Syamlal-Obrien drag submodel is 1.64% accurate than in comparison to both the Laminar model and RANS model with Gidaspow drag sub-model. The
ACCEPTED MANUSCRIPT largest difference between predicted gas volume fractions and experimental measurement appeared in the distance interval between 0.15 and 0.28 meters. This can be attributed to the fact that the choice of appropriate multiphase model or the drag sub-model along might probably not improve the numerical results. Some other parameters such as frictional viscosity sub- model, specularity coefficient, and wall roughness are not properly modeled in the bed. The influence of frictional viscosity is depicted using both grids. The RANS model without
T
frictional viscosity demonstrates an averaged excessive prediction, 8.16% across the full
IP
cross-section, whereas the model with frictional viscosity shows an averaged deviation of 5.35%.
CR
The stress that is generated by friction between the particles can be contributed to the solids shear viscosity. Consequently, higher energy loss and lower gas volume fraction appeared due to the
Discussion
AN
5
US
high viscous solid particles.
Before moving to the conclusion of the present study, it is important to discuss the appropriateness
M
of 2D simulations to validate different numerical models and sub- models. The number of
ED
publications reporting the results of the 2D modeling of fluidized beds, e.g. see works [27, 19, 38, 39] and a list of references within these publications, tremendously overweights a limited number of the 3D studies. The reason is well understood: the 3D simulations are associated with significant
PT
computational demands and time. Xie et. al. [28, 29] showed that in some cases the similarities between 2D and 3D simulations can be observed for bubbling regimes with low gas velocity. Hu
CE
et. al. [39] used the 2D simulations to validate their EMMS drag model. Excellent agreement with the experimental data was demonstrated. However, analysis of the results of our present work
AC
shows that the 2D model does not produce physically meaningful results in terms of void fraction distribution. This discrepancy is attributed to neglecting the third dimension in 2D simulations. The results of the 3D simulations clearly show that wall friction in the third direction plays a significant role in the size of bubbles and in the spatial distribution of the volume fraction of particles inside the fluidization zone. We show that a reasonable set up of model parameters (without any sophisticated new drag models) produces very good agreement with the experimental data when 3D simulations are used. We emphasize that a validation of new sub- models against experimental data using 2D simulations does not guarantee that such sub- models work well in 3D
ACCEPTED MANUSCRIPT simulations giving physically realistic results. Moreover, a certain combination of numerical parameters for the 2D simulations can result in acceptable agreement with the experimental data which could be a misleading finding. We suggest that 3D models are used to validate the newly-developed sub- models and to select the model parameters. Only then the validated sub-models and an outlined set of parameters can be used for 2D sensitivity analyses. Finally, we want to discuss the role of turbulence and its modeling on the results presented
T
in this work. We used a well accepted and well used k - dispersed turbulence model, which
IP
showed a reasonable agreement with the experimental studies [18]. The analysis of literature
CR
carried out in section Introduction showed that many 2D and 3D CFD simulations of fluidized beds using laminar flow model, e.g., see [23, 21, 22], showed reasonable agreement with the
US
experimental data published in the literature. Our simulations revealed that the turbulence inside the bed is significantly weaker than in the gas phase outside of the bed. However, there are
Conclusions
M
6
AN
numerous works devoted to fundamental studies of the turbulence in fluidized beds [40, 41].
ED
In the present study, the numerical simulations of a fluidized bed filled with the Geldart Group B particles were performed using the CFD commercial software, FLUENT 16.2. The two- and three-dimensional approaches were considered. A gas-solid phase Eulerian-Eulerian based model
PT
was used to replicate the experimental of Taghipour[14]. The influence of the model dimension (2D & 3D), the flow regimes model (laminar & turbulent), the model parameters (the restitution
CE
coefficient responsible for particle-particle interaction and the specularity coefficient responsible for the wall-particle interaction) and the computational girds were studied. Following are the main
AC
conclusions that can be drawn from this study: The 3D and 2D approaches showed a deviation of a time-averaged void fraction within [3.85-8.54]% and [3.97-13.08]%, respectively.
The results of the 3D simulations clearly illustrate that the finite size of the domain where fluidization occurs plays a significant role in the size of bubbles and in spatial distribution of the volume fraction of particles inside the fluidization zone.
We showed that the 2D approach does not give a good agreement with the experimental data. However, some combination of the grid resolution, the specularity and restitution
ACCEPTED MANUSCRIPT coefficients can give results close to the experimental points.
It was shown that the turbulence does not play a significant role inside the fluidized bed. However, the use of model provides slightly better agreement with experimental data in comparison with results given by a laminar model.
Taking into account the frictional viscosity provides better agreement with the
The specularity coefficient has a significant influence on the voidage distribution inside the
IP
T
experimental data.
bed in comparison to the restitution coefficient under conditions used in the experiment by
CR
Taghipour[14].
Christopher Higman and Maarten Van der Burgt. Gasification. Gulf professional
AN
[1]
US
References publishing, 2011. [2]
P. Roache, K. Ghia, and F. White. Editorial policy statement on the control of numerical
P. Roache. Verification and Validation in Computational Science and Engineering.
ED
[3]
M
accuracy. ASME J. Fluid Eng, 108:1, 1986.
Hermosa Publishers, Albuquerque, NM, 1998. P. Roache. Perspective: Validation{what does it mean? ASME J. Fluid Eng, 131:034503,
PT
[4]
2008.
Ansys Fluent. 12.0 theory guide. Ansys Inc, 5, 2009.
[6]
CFX Ansys. Solver theory guide. Ansys CFX Release, 11:1996{2006, 2006.
[7]
USER CD-adapco. Guide: Star-ccm+, 2009.
[8]
CW Hirt and B Nichols. Flow-3D user's manual. Flow Science Inc, 1988.
[9]
Alexander Stroh, Falah Alobaid, Max Thomas Hasenzahl, Jochen Hilz, Jochen Ströhle,
AC
CE
[5]
and Bernd Epple. Comparison of three different CFD methods for dense fluidized beds and validation by a cold flow experiment. Particuology, 2016. [10]
Chanchal Loha, Himadri Chattopadhyay, and Pradip K Chatterjee. Euler-Euler CFD modeling of fluidized bed: Influence of specularity coefficient on hydrodynamic behavior. Particuology, 11(6):673–680, 2013.
ACCEPTED MANUSCRIPT [11]
Chanchal Loha, Himadri Chattopadhyay, and Pradip K Chatterjee. Effect of coefficient of restitution in euler–euler CFD simulation of fluidized-bed hydrodynamics. Particuology, 15: 170–177, 2014.
[12]
David G Schaeffer. Instability in the evolution equations describing incompressible granular flow. Journal of differential equations, 66(1):19–50, 1987. Peter McGinnis. Biomechanics of sport and exercise. Human Kinetics, 2013.
[14]
Fariborz Taghipour, Naoko Ellis, and Clayton Wong. Experimental and computational
T
[13]
IP
study of gas–solid fluidized bed hydrodynamics. Chemical Engineering Science,
CR
60(24):6857–6867, 2005.
Madhava Syamlal and Thomas J O'Brien. Computer simulation of bubbles in a fluidized
US
[15]
bed. In AIChE Symp. Ser, volume 85, pages 22–31, 1989. Dimitri Gidaspow. Multiphase flow and fluidization: continuum and kinetic theory descriptions. Academic press, 1994.
CY Wen and Y Yu. Mechanics of fluidization. In Chem. Eng. Prog. Symp. Ser., volume 62,
M
[17]
page 100, 2013.
Naser Almohammed, Falah Alobaid, Michael Breuer, and Bernd Epple. A comparative
ED
[18]
AN
[16]
study on the influence of the gas flow rate on the hydrodynamics of a gas–solid spouted
PT
fluidized bed using Euler–Euler and Euler–Lagrange/DEM models. Powder Technology, 264:343–364, 2014.
J.T. Cornelissen, F. Taghipour, R. Escudié, N. Ellis, and J.R. Grace. CFD modelling of a
CE
[19]
liquid–solid fluidized bed. Chemical Engineering Science, 62(22):6334–6348, 2007. Baosheng Jin, Xiaofang Wang, Wenqi Zhong, He Tao, Bing Ren, and Rui Xiao. Modeling
AC
[20]
on high-flux circulating fluidized bed with Geldart group B particles by kinetic theory of granular flow. Energy & fuels, 24:3159–3172, 2010. [21]
A Bakshi, C Altantzis, RB Bates, and AF Ghoniem. Eulerian–Eulerian simulation of dense solid–gas cylindrical fluidized beds: Impact of wall boundary condition and drag model on fluidization. Powder Technology, 277:47–62, 2015.
[22]
C. Altantzis, R.B. Bates, and A.F. Ghoniem. 3D Eulerian modeling of thin rectangular gas-solid fluidized beds: Estimation of the specularity coefficient and its effects on bubbling dynamics and circulation times. Powder Technology, 270:256–270, 2015.
ACCEPTED MANUSCRIPT [23]
Tingwen Li, John Grace, and Xiaotao Bi. Study of wall boundary condition in numerical simulations of bubbling fluidized beds. Powder Technology, 203:447–457, 2010.
[24]
Lei Kong, Chao Zhang, and Jesse Zhu. Evaluation of the effect of wall boundary conditions on numericalsimulations of circulating fluidized beds. Particuology, 13:114–123, 2014.
[25]
Wenqi Zhong, Aibing Yu, Guanwen Zhou, Jun Xie, and Hao Zhang. CFD simulation of
T
dense particulate reaction system: Approaches, recent advances and applications.
Martin Anton van der Hoef, M van Sint Annaland, NG Deen, and JAM Kuipers. Numerical
CR
[26]
IP
Chemical Engineering Science, 140:16–43, 2016.
simulation of dense gas-solid fluidized beds: a multiscale modeling strategy. Annu. Rev.
[27]
US
Fluid Mech., 40:47–70, 2008.
N. Yang, W. Wang, W. Ge, and J. Li. CFD simulation of concurrent-up gas-solid flow in
AN
circulating fluidized beds with structure-dependent drag coefficient. Chemical Engineering Journal, 96:71–80, 2003.
N. Xie, F. Battaglia, and S. Pannala. Effects of using two- versus three-dimensional
M
[28]
computational modeling of fluidized beds part i, hydrodynamics. Powder Technology,
ED
182:1–13, 2008.
N. Xie, F. Battaglia, and S. Pannala. Effects of using two- versus three-dimensional
PT
[29]
computational modeling of fluidized beds: Part ii, budget analysis. Powder Technology,
[30]
CE
182:14–24, 2008.
C Simonin and PL Viollet. Predictions of an oxygen droplet pulverization in a
AC
compressible subsonic coflowing hydrogen flow. Numerical Methods for Multiphase Flows, FED91, pages 65–82, 1990. [31]
JO Hinze. Turbulence, mcgraw-hill publishing co. New York, NY, 1975.
[32]
Sebastian Zimmermann and Fariborz Taghipour. CFD modeling of the hydrodynamics and reaction kinetics of fcc fluidized-bed reactors. Industrial & engineering chemistry research, 44 (26):9818–9827, 2005.
[33]
Ling Zhou, Lingjie Zhang, Ling Bai, Weidong Shi, Wei Li, Chuan Wang, and Ramesh Agarwal. Experimental study and transient CFD/DEM simulation in a fluidized bed based on different drag models. RSC Advances, 7(21):12764–12774, 2017.
ACCEPTED MANUSCRIPT [34]
Brian P Leonard. A stable and accurate convective modelling procedure based on quadratic upstream interpolation. Computer methods in applied mechanics and engineering, 19(1):59–98, 1979.
[35]
S Muzaferija, M Peric, P Sames, and T Schellin. A two-fluid navier-stokes solver to simulate water entry. In Proceedings of the 22nd symposium on naval hydrodynamics, Washington, DC, pages 277{289, 1998. Henry Scheffe. The relation of control charts to analysis of variance and chi-square tests.
T
[36]
P. Nikrityuk, M. Ungarish, K. Eckert, and R. Grundmann. Spin-up of a liquid metal flow
CR
[37]
IP
Journal of the American Statistical Association, 42(239):425–431, 1947.
driven by a rotating magnetic field in a finite cylinder: A numerical and an analytical study.
[38]
US
Phys. Fluids, 17:067101, 2005.
Mehran Ehsani, Salman Movahedirad, and Shahrokh Shahhosseini. The effect of particle
AN
properties on the heat transfer characteristics of a liquid–solid fluidized bed heat exchanger. International Journal of Thermal Sciences, 102:111–121, 2016. S. Hu, X. Liu, N. Zhang, J. Li, W. Ge, and W. Wang. Quantifying cluster dynamics to
M
[39]
improve EMMS drag law and radial heterogeneity description in coupling with gas-solid
[40]
ED
two-fluid method. Chemical Engineering Journal, 307:326–338, 2017. Jesse Capecelatro, Olivier Desjardins, and Rodney O. Fox. Strongly coupled fluid-particle
PT
flows in vertical channels. i. reynoldsaveraged two-phase turbulence statistics. Physics of Fluids, 28: 033306, 2016.
S. Schneiderbauer. Validation study on spatially averaged two-fluid model for gas-solid
CE
[41]
flows: Ii. application to risers and fluidized beds. AIChE Journal, 64:1606–1617, 2018. Adnan Almuttahar and Fariborz Taghipour. Computational fluid dynamics of a circulating
AC
[42]
fluidized bed under various fluidization conditions. Chemical Engineering Science, 63(6):1696–1709, 2008. [43]
Leina Hua, Junwu Wang, and Jinghai Li. CFD simulation of solids residence time distribution in a cfb riser. Chemical Engineering Science, 117:264–282, 2014.
[44]
Tingwen Li, Aytekin Gel, Sreekanth Pannala, Mehrdad Shahnam, and Madhava Syamlal. CFD simulations of circulating fluidized bed risers, part i: Grid study. Powder Technology, 265:2–12, 2014.
ACCEPTED MANUSCRIPT [45]
Thanh DB Nguyen, Myung Won Seo, Young-Il Lim, Byung-Ho Song, and Sang-Done Kim. CFD simulation with experiments in a dual circulating fluidized bed gasifier. Computers & Chemical Engineering, 36:48–56, 2012.
[46]
Regis Andreux, Geoffrey Petit, Mehrdji Hemati, and Olivier Simonin. Hydrodynamic and solid residence time distribution in a circulating fluidized bed: experimental and 3D computational study. Chemical Engineering and Processing: Process Intensification,
CKK Lun, S Br Savage, DJ Jeffrey, and N Chepurniy. Kinetic theories for granular flow:
IP
[47]
T
47(3):463–473, 2008.
Journal of fluid mechanics, 140:223–256, 1984.
Paul C Johnson and Roy Jackson. Frictional–collisional constitutive relations for granular
US
[48]
CR
inelastic particles in couette flow and slightly inelastic particles in a general flowfield.
materials, with application to plane shearing. Journal of fluid Mechanics, 176:67–93,
[49]
AN
1987.
Jianmin Ding and Dimitri Gidaspow. A bubbling fluidization model using kinetic theory of
M
granular flow. AIChE journal, 36(4):523–538, 1990.
ED
Table 1: Summary of published studies. The ratios x / d p , y / d p , and z / d p represent the number of particles along each grid edge; E-E = Eulerian-Eulerian; E-L = Eulerian-Lagrangian; E-L-S = Eulerian-Lagrangian-stochastic; E-L-D = Eulerian-Lagrangian-deterministic x / d p , y / d p , z / d p
E-E, E-L-S, E-L-D
0.75, 0.36, 0.93
Gray Almohammed [18]
E-E, E-L
0.75, 0.36, 1.85
Gray Ehsani [38]
E-E
0.33, 0.33
Cornelissen [19]
E-E
1.87, 1.77
Almuttahar [42]
E-E
14.5, 283
Loha[11]
E-E
3.66, 24.3
Bakshi [21]
E-E
25.1, 17.3
AC
CE
Gray Stroh [9]
PT
Model
Authors
8.30, 5.0 20.6, 17.3 Hua [43]
E-E
65.5, 65.5, 453
Li [44]
E-E
45.7, 94.1
ACCEPTED MANUSCRIPT 11.4, 23.5 22.8, 47.1 20, 20 14.1, 14.1 10, 9.98 E-E
20, 20
Andreux [46]
E-E
157.1, 412.1, 157.1
IP
T
Nguyen [45]
CR
Table 2: List of different grid resolutions used in simulations of fluidized beds. Number of cells
x / d p , y / d p , z / d p
D-Grid1
56 300 = 16800
18.18, 18.18
D-Grid2
112 600 = 67200
D
56 300 5 = 84000 18.18, 18.18, 18.18
US
Grid
AN
9.09, 9.09
M
Table 3: List of different models and schemes used in simulations of fluidized beds. Name
Model/ Scheme Name Eulerian-Eulerian [5]
ED
Multiphase Flow
Implicit Scheme [5]
Viscous Model
Laminar & RANS- k - model (dispersed) [5]
Drag Model
PT
Volume Fraction Parameters
Granular Temperature
Syamlal-O’Brien & Gidaspow [15], [16], [5] Algebraic [5] Syamlal-O’Brien [15], [5]
Granular Bulk Viscosity
Lun et al. [47], [5]
Frictional Viscosity
Johnson-et-al [48], [5]
Frictional Pressure (pascal)
Based-ktgf [49], [5]
Solid Pressure
Syamlal-O’Brien [15], [5]
Radial Distribution
Syamlal-O’Brien [15], [5]
Pressure-Velocity Coupling
Coupled Scheme [5]
Spatial Discretization-Gradient
Least Squares Cell Based [5]
Spatial Discretization-Momentum
QUICK [34]
Spatial Discretization-Volume
Modified HRIC [35]
AC
CE
Granular Viscosity
Fraction
ACCEPTED MANUSCRIPT k - 2nd order Upwind [5]
Transient Formulation
First Order Implicit
Flow
velocity U,
model
Drag force
m/s 3D-Grid
0.38
Laminar
1 0.38
3D-Grid
0.38
1 4
3D-Grid
0.38
1 5
3D-Grid
0.46
1 3D-Grid
0.46
0.1
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
0.9
0.95
Syamlal-O’
0.9
0.9
Johnson-et-a
0.1
None
0.1
None
1
Johnson-et-a
l
l 0.1
Johnson-et-a l
0.9
0.5
None
0.9
0.1
None
0.9
1
Johnson-et-a
Brien
RANS
Syamlal-O’
k -
Brien
2D-Grid 0.38
RANS
Gidaspow
0.9
0.1
None
1
k -
0.38
Laminar(H
Syamlal-O’
0.9
0.1
None
RS)
Brien
Laminar(H
Syamlal-O’
0.9
0.1
Johnson-et-a
RS)
Brien
Laminar(U
Syamlal-O’
DS)
Brien
RANS
Syamlal-O’
k -
Brien
0.46
2D-Grid
2D-Grid
0.38
2D-Grid
0.38
2D-Grid 1
0.38
ED
PT
3D-Grid
1 13
0.9
k -
0.46
1 12
K
Syamlal-O’
1 11
ess
RANS
3D-Grid
AC
10
viscosity
Brien
1 9
coefficient,
k -
1 8
RANS
RANS
1 7
coefficient,
CE
6
Frictional
US
3D-Grid 1
3
Specularity
Brien
AN
2
Syamlal-O’
M
1
Restitution
T
Inlet
IP
Grid
CR
Case
Spatial Discretization-Turbulence
l
l 0.9
0.1
None
0.9
0.1
None
ACCEPTED MANUSCRIPT
0.38
1 16
2D-Grid
0.38
1 17
2D-Grid
0.38
1 18
2D-Grid
0.38
1 19
2D-Grid
0.38
1 20
2D-Grid
0.38
1 21
2D-Grid
0.38
1 22
2D-Grid
0.38
0.38
2D-Grid
2D-Grid 2
27
2D-Grid
2D-Grid
2D-Grid 1
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
Syamlal-O’
0.9
0.9
None
0.99
0.1
None
0.95
0.1
None
0.9
0.9
0.99
0.99
0.01
None
0.1
Johnson-et-a l
0.01
Johnson-et-a l
0.9
Johnson-et-a l
0.9
1
Johnson-et-a
k -
Brien
RANS
Gidaspow
0.9
0.1
None
Syamlal-O’
0.9
0.1
None
0.9
0.1
None
0.9
1
Johnson-et-a
Laminar
l
Brien Syamlal-O’
k -
Brien
0.38
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
0.38
0.38
2 29
Syamlal-O’
RANS
0.38
2 28
RANS
PT
0.38
AC
2 26
Brien
CE
2D-Grid 2
25
k -
None
k -
2 24
Syamlal-O’
ED
2D-Grid
RANS
RANS
1 23
Brien
0.5
T
2D-Grid
k -
0.9
IP
15
Syamlal-O’
CR
1
RANS
US
0.38
AN
2D-Grid
M
14
0.46
l 0.9
0.01
Johnson-et-a l
0.9
0.1
Johnson-et-a l
0.9
0.1
Johnson-et-a l
ACCEPTED MANUSCRIPT 30
2D-Grid
0.46
1 31
2D-Grid
0.46
2
RANS
Syamlal-O’
k -
Brien
RANS
Syamlal-O’
k -
Brien
0.9
0.5
None
0.9
0.5
None
FIGURES
IP
7
T
Table 4: Parameters of simulation cases performed in this study.
CR
List of Figures
Figure 1: Graphical explanation of (a) – the restitution coefficient ess and (b) – the specularity
US
coefficient K .
AN
Figure 2: Schematic of a fluidized bed reactor, adapted from [14]
M
Figure 3: Simulation errors for 3D and 2D models.
ED
Figure 4: 3D-RANS simulations. Snapshots at different time instances of (a) – solid volume fraction and (b) – turbulent viscosity ratio. Simulation parameters: drag function is modelled by
PT
Syamlal–O’Brien sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1
CE
; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
Figure 5: 3D-RANS simulations. Time-averaged solid volume fraction. Simulation parameters:
AC
drag function is modelled by Syamlal–O’Brien sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
Figure 6: Time history of the volume-averaged solid velocity ( U s ) predicted numerically using 2D RANS and 3D RANS simulations with the restitution coefficient of ess = 0.9 , the drag function is modelled by Syamlal–O’Brien sub-model; inlet velocity U = 0.38 m/s
Figure 7: 3D-RANS simulations. Experimental [14] and numerical profiles of time-averaged void
ACCEPTED MANUSCRIPT (gas phase) along x axis at z = 0.2 m. Simulation parameters: the drag function is modelled by Syamlal–O’Brien sub-model.(a) – U = 0.38 m/s; (b) – U = 0.46 m/s; (c) - comparison between RANS and laminar models for U = 0.38 m/s and K = 0.1 , ess = 0.9 . Figure 8: 2D simulations on Grid1, 56 300 . Solid volume fraction at different time instances
T
obtained as a results of unsteady (a) – Laminar and (b) – RANS simulations. Simulation
IP
parameters: drag function is modelled by Syamlal-O’Brien sub-model; no frictional viscosity
CR
sub-model; specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity
U = 0.38 m/s.
US
Figure 9: 2D simulations on Grid2, 112 600 . Solid volume fraction at different time instances obtained as a results of unsteady (a) – Laminar and (b) – RANS simulations. Simulation
AN
parameters: drag function is modelled by Syamlal-O’Brien sub-model; no frictional viscosity sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1 ; restitution
M
coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
ED
Figure 10: 2D simulation on Grid1, 56 300 . Time-averaged solid volume fraction calculated using unsteady (a1) – Laminar and (a2) – RANS E-E model. Time-averaged gas velocity vector
PT
calculated using unsteady (b1) – Laminar and (b2) – RANS E-E model. Simulation parameters:
CE
drag function is modelled by Syamlal-O’Brien sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
AC
Figure 11: 2D simulation on Grid2, 112 600 . Time-averaged solid volume fraction calculated using unsteady (a1) – Laminar and (a2) – RANS E-E model. Time-averaged gas velocity vector calculated using unsteady (b1) – Laminar and (b2) – RANS E-E model. Case parameters: drag function is modelled by Syamlal-O’Brien sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s. Figure 12: 2D-RANS simulations. Turbulent viscosity ratio ( t / 0 ) at different time instances. Simulation parameters: drag function is modelled using Syamlal O’Brien sub- model; no frictional
ACCEPTED MANUSCRIPT viscosity sub-model; specularity coefficient: K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s. (a) – Grid1 and (b) – Grid2. Figure 13: Time-averaged gas velocity ( U g ) as a function of height at x = 0.14 m (a) 2D model with frictional viscosity and (b) – 3D model. Simulation parameters: the drag function is
T
modelled by Syamlal O’Brien sub-model; the frictional visocisty is modelled by Johnson-et-al
IP
sub-model; specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity
CR
U = 0.38 m/s.
US
Figure 14: Volume-averaged solid velocity ( U s ) as a function of time for (a) – 2D model with frictional viscosity and (b) – 3D model. Simulation parameters: the drag function is modelled by
AN
Syamlal O’Brien sub-model; the frictional visocisty is modelled by Johnson-et-al sub-model;
M
specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
Figure 15: 2D-RANS simulations. Experimental [14] and numerical profiles of time-averaged
ED
void (gas phase) along x axis at z = 0.2 m for different values of (a) – restitution coefficient
ess and (b) – specularity coefficient K . Simulation parameters: the drag function is modelled by
PT
Syamlal O’Brien sub-model; no frictional viscosity sub-model; specularity coefficient K = 0.1 ;
CE
restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s.
Figure 16: Time history of the volume-averaged solid velocity ( U s ) predicted numerically using
AC
3D RANS simulations. Simulation parameters: the drag function is modelled by Syamlal O’Brien sub-model; no frictional viscosity sub-model; restitution coefficient ess = 0.9 ; inlet velocity
U = 0.38 m/s. Figure 17: 2D simulations. Experimental [14] and numerical profiles of time-averaged void fraction at z = 0.2 m. Simulation parameters: specularity coefficient K = 0.1 ; restitution coefficient ess = 0.9 ; inlet velocity U = 0.38 m/s (a) – Grid1, and (b) – Grid2.
ACCEPTED MANUSCRIPT
AC
CE
PT
ED
M
AN
US
CR
IP
T
Highlights Simulations of fluidized bed were performed in ANSYS FLUENT using Euler-Euler model The results were validated against reference experimental data 3D simulations give more physically realistic results as opposed to 2D simulations Turbulent flow model should be used to simulate gasification and combustion The choice of simulation parameters significantly affects numerical predictions
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14
Figure 15
Figure 16
Figure 17