PDF

Dispersion of Heavy Particles in a Turbulent Channel Flow
Introduction
In this benchmark model, solid particles are released in a fully developed turbulent channel flow. The total force acting on particles in a fluid comprises a large number of physical phenomena, including but not limited to the drag force, gravity force, buoyancy force, pressure gradient force, added mass effect, lift force, and Brownian force. In this example, the drag force is assumed to be the dominant factor in determining the particle trajectories.
If the flow in the turbulent channel flow is modeled using the Reynolds-averaged Navier-Stokes (RANS) equations, then the fluid velocity is treated as the sum of a deterministic mean flow term and a random velocity perturbation that represents the eddies.
The advantage of the RANS equations over Direct Numerical Simulation (DNS) is that the RANS equations can be solved without requiring a mesh that is sufficiently fine to resolve all of the eddies in the flow, which may be impractically computationally expensive. Furthermore, by treating the chaotic aspect of turbulent flow in a statistical or time-averaged sense, the RANS equations permit stationary solutions to turbulent flow problems, while in reality the flow field is constantly evolving as eddies of all sizes are created, transported, and destroyed.
This example provides an overview of the essential considerations when modeling the motion of particles in a turbulent channel flow. Among the factors considered in this example are:
While this example couples the Particle Tracing for Fluid Flow interface to the fluid velocity field and turbulent kinetic energy based on a RANS model, the resulting particle distributions are compared to DNS data compiled by Marchioli et al. (Ref. 1), which show reasonably good agreement. As in Ref. 1, the particles are observed to cluster near the channel wall for an intermediate range of Stokes numbers where they have sufficient inertia to cross some eddies in the flow but not enough to consistently be reflected at the channel wall.
Model Definition
The model geometry is a 2D vertical channel containing air. The model parameters are the same as those used in the DNS simulation in Ref. 1:
ν = 15.7 × 10-6 m2/s is the air kinematic viscosity,
ρ = 1.3 kg/m3 is the air density,
h = 0.02 m is the channel half-width,
va = 1.65 m/s is the average fluid velocity, and
ρp = 769ρ is the particle density.
These parameters yield a value of Re = vah/ν = 2100 for the Reynolds number of the flow. The authors predict a shear velocity of uτ = 0.11775 m/s for a shear Reynolds number of approximately Reτ = uτh/ν = 150.
Turbulent Flow
This example solves the Reynolds-averaged Navier-Stokes (RANS) equations and uses the standard k-ε turbulence model (Ref. 2), one of the most frequently used turbulence models in computational fluid dynamics. The k-ε model introduces dependent variables and transport equations for two new quantities:
The turbulent kinetic energy k (SI unit: m2/s2) represents the energy per unit mass associated with eddies in the flow.
The turbulent dissipation rate ε (SI unit: m2/s3) indicates the rate at which turbulent kinetic energy in the eddies is converted to thermal energy.
These new dependent variables provide insight into the size and lifetime of eddies in the flow:
The ratio k has units of time. The average eddy lifetime is often estimated by multiplying this ratio by a dimensionless constant of order unity.
The ratio k3/2 has units of length and indicates the length scale of the largest eddies in the flow.
In addition to being a very common turbulence model in its own right, the k-ε model is relatively easy to combine with Lagrangian particle tracking methods because the turbulence variables immediately give estimates of the amplitude of the velocity perturbations due to turbulent eddies (proportional to ) and the average eddy lifetime (proportional to k). For a stationary, incompressible flow, the transport equations solved when using the k-ε turbulence model are
where the dependent variables are the fluid velocity u (SI unit: m/s), pressure p (SI unit: Pa), and the aforementioned transport variables k and ε.
The dimensionless constants in these equations have the following default values:
For more details on the different turbulence models that are available, see the Single Phase Flow Interfaces chapter in the CFD Module User’s Guide. The k-ε model is usually a convenient choice when coupling the turbulent flow to a particle tracing simulation because of its relatively low memory requirements and fast convergence, and because it directly allocates degrees of freedom for k and ε, which are necessary to accurately model turbulent dispersion of the particles.
Lagrangian Particle Tracking
Marchioli et al. (Ref. 1) provide a wide range of DNS results for simulations with and without the lift and gravity forces. In this example, only the drag force is considered.
Because the particle density is several orders of magnitude greater than the air density, the buoyant force and added-mass effect can safely be neglected. The particles are also large enough that the Brownian Force can be ignored.
The equation of motion for each particle is therefore
where
mp (SI unit: kg) is the particle mass,
q (SI unit: m) is the particle position, and
FD (SI unit: N) is the drag force.
In general, the drag force is defined as
(1)
where u (SI unit: m/s) is the fluid velocity at the particle’s position and v (SI unit: m/s) is the particle velocity. The particle relaxation time τp (SI unit: s) is defined as
(2)
where
ρp (SI unit: kg/m3) is the particle density,
dp (SI unit: m) is the particle diameter, and
CD (dimensionless) is the drag coefficient.
The choice of drag law, which determines the appropriate definition of CD, depends on the relative Reynolds number Rer (dimensionless) of the particle in the fluid. For a spherical particle the relative Reynolds number is
For the Stokes drag law is applicable,
so that Equation 2 reduces to
Particle Tracing with Turbulent Dispersion
In the Reynolds-Averaged Navier-Stokes (RANS) approach, the turbulent eddies are only solved for in a statistical sense, by estimating their energy and dissipation rate. A Direct Numerical Simulation (DNS) could resolve the individual eddies but is often too memory-intensive and time-consuming for many practical applications. Therefore, in Equation 1 the fluid velocity u is not given deterministically. Instead, the fluid velocity is treated as a linear combination of a deterministic part based on the mean flow (which is the field u solved for by the RANS equations) and a turbulent perturbation term Δu,
The amplitude and direction of the velocity perturbation Δu is derived from the turbulence variables k and ε that are solved for by the k-ε turbulence model.
The Particle Tracing for Fluid Flow interface supports two different formulations for the turbulent dispersion term:
The Discrete Random Walk (DRW) model is similar to the modified Eddy Interaction Model of Gosman and Ioannides (Ref. 3). In the DRW model, unique velocity perturbations are sampled and added to the mean velocity field at discrete times based on the estimated eddy lifetime.
The Continuous Random Walk (CRW) model, or Continuous Filter White Noise (CFWN) model (Ref. 4). In the CRW model, the velocity perturbations are integrated over time. The eddy velocity and lifetime are built into the time derivatives of the evolving velocity perturbation components.
In this example, the CRW model will be used. A brief overview of this model is given below. For more comprehensive details on both turbulent dispersion models, see the Particle Tracing for Fluid Flow chapter of the Particle Tracing Module User’s Guide.
The classical Langevin-equation model for homogeneous isotropic stationary turbulence (HIST) is (Ref. 5)
(3)
where the subscript indicates a component of the fluid velocity field. In isotropic turbulence, the RMS fluid velocity fluctuation in any direction is equal,
Because the turbulence is isotropic, the velocity perturbations can be aligned with any orthonormal coordinate system. In the Particle Tracing for Fluid Flow interface they are simply aligned with the Cartesian coordinates.
Inhomogeneous and Anisotropic Turbulence
In wall-bounded flows the assumption of homogeneous isotropic turbulence no longer applies. The turbulence becomes inhomogeneous because the turbulent kinetic energy is heavily damped close to the walls, giving it a nonzero gradient in this region. The turbulence is anisotropic because the eddies in the region close to the wall are not equally likely to point in any direction; the instantaneous velocity component normal to the wall is more heavily damped than those in the streamwise and spanwise directions.
In the Particle Tracing for Fluid Flow interface, anisotropic and inhomogeneous turbulent dispersion are modeled by selecting the Include anisotropic turbulence in boundary layers check box in the settings for the Drag Force node. This check box is only available when the Continuous Random Walk model for turbulent dispersion is used. When this check box is selected, corrections for inhomogeneous and anisotropic turbulence are applied in the region y+ < 100, where y+ (dimensionless) is the wall distance in viscous units,
where x2 (SI unit: m) is the normal distance to the nearest wall, ν (SI unit: m2/s) is the kinematic viscosity of the fluid, and uτ (SI unit: m/s) is the friction velocity,
where τw (SI unit: N/m2) is the wall shear-stress. The friction velocity is usually defined on the Wall boundaries by one of the turbulent flow interfaces. In the region of inhomogeneous, anisotropic turbulence, the components of the turbulent velocity perturbation are defined in the coordinate system given by
The normalized Langevin equations in these directions are
To account for the anisotropy of the flow, the following definitions of the σi terms are given. These expressions are taken from Dehbi (Ref. 5), wherein they are attributed to DNS fits of channel flow as described by Dreeben and Pope (Ref. 6).
In the boundary layer, the Lagrangian time scale τi from Equation 3 is approximately the same in all directions:
Kallio and Reeks Ref. 7 give the following polynomial fit for the time scale:
where ν (SI unit: m2/s) is the fluid kinematic viscosity. Away from the wall, the Lagrangian time scale is simply
where CL is a dimensionless constant.
Effect of Particle Inertia
To investigate the effect of particle inertia on anisotropic turbulent dispersion in the channel, a Parametric Sweep is performed over the Stokes number St (dimensionless) of the particle:
where τp (SI unit: s) is the particle relaxation time,
Marchioli et al. (Ref. 1) use six values of the Stokes number: 0.2, 1, 5, 15, 25, and 125. For a given value of St, the corresponding particle diameter is
Results and Discussion
The results of the turbulent flow simulation are shown in Figure 1 and Figure 2. The channel has an extremely high aspect ratio, so to better visualize the geometry, Automatic has been selected from the View scale list in the Axis settings. This allows the coordinate axes to be scaled independently of each other so that the plot fits the Graphics window.
Figure 1 shows the fluid velocity magnitude and velocity streamlines. Figure 2 shows the ratio of turbulent kinetic energy to turbulent dissipation rate in the modeling domain. This is greatest near the left boundary, which is the axis of symmetry. It decreases near the wall, suggesting that eddy lifetimes are much shorter there. It is useful to plot this ratio before proceeding with the particle tracing simulation because it shows the minimum resolution in time needed to accurately capture the particle-eddy interactions.
The particle trajectories for the greatest simulated value of the Stokes number, St = 125, are shown in Figure 3. The particles were released uniformly along the cross section halfway down the channel, to prevent them from hitting any Inlet or Outlet boundaries as a result of the turbulent diffusion. For intermediate values of the Stokes number, the particles cluster near the wall, as shown by the histograms in Figure 4. At very low Stokes number, St = 0.2, the effect of anisotropic turbulence on the number density of particles is less pronounced because such particles don’t have enough inertia to frequently cross the eddies. At very high Stokes number, St = 125, the effect is less pronounced compared to intermediate values because the inertia of such particles is so high that they often reflect off the wall and back into the bulk medium.
Figure 1: Fluid velocity in the channel. Velocity streamlines are shown as white lines.
Figure 2: Turbulence time scale, or ratio of turbulent kinetic energy to turbulent dissipation rate, in the channel.
Figure 3: Particle trajectories in the channel.
Figure 4: Comparison of histograms of particle position for different values of the Stokes number.
References
1. C. Marchioli, M. Picciotto, and A. Soldati, “Influence of gravity and lift on particle velocity statistics and transfer rates in turbulent vertical channel flow”, International Journal of Multiphase Flow, vol. 33, no. 3, pp. 227–251, 2007.
2. D.C. Wilcox, Turbulence Modeling for CFD, 2nd ed., DCW Industries, 1998.
3. A. D. Gosman and E. Ioannides, “Aspects of computer simulation of liquid-fueled combustors”, Journal of Energy, vol. 7, no. 6, pp. 482–490, 1983.
4. L. Tian and G. Ahmadi, “Particle deposition in turbulent duct flows–comparisons of different model predictions”, Aerosol Science, vol. 38, 2007, pp. 377–397.
5. A. Dehbi, “Turbulent particle dispersion in arbitrary wall-bounded geometries: A coupled CFD-Langevin-equation based approach”, International Journal of Multiphase Flow, vol. 34, no. 9, pp. 819–828, 2008.
6. T. D. Dreeben and S. B. Pope, “Probability density function and Reynolds-stress modeling of near-wall turbulent flows”, Physics of Fluids, vol. 9, no. 1, pp. 154–163,1997.
7. G. A. Kallio and M. W. Reeks, “A numerical simulation of particle deposition in turbulent boundary layers”, International Journal of Multiphase Flow, vol. 15, no. 3, pp. 433–446, 1989.
Application Library path: Particle_Tracing_Module/Fluid_Flow/flow_channel_turbulent_dispersion
Modeling Instructions
From the File menu, choose New.
New
In the New window, click  Model Wizard.
Model Wizard
1
In the Model Wizard window, click  2D.
2
In the Select Physics tree, select Fluid Flow>Single-Phase Flow>Turbulent Flow>Turbulent Flow, k-ε (spf).
3
Click Add.
4
Click  Study.
5
In the Select Study tree, select General Studies>Stationary.
6
Global Definitions
Parameters 1
Load the model parameters from a file.
1
In the Model Builder window, under Global Definitions click Parameters 1.
2
In the Settings window for Parameters, locate the Parameters section.
3
Click  Load from File.
4
Geometry 1
Rectangle 1 (r1)
1
In the Geometry toolbar, click  Rectangle.
2
In the Settings window for Rectangle, locate the Size and Shape section.
3
In the Width text field, type halfWidth.
4
In the Height text field, type height.
5
Click  Build All Objects.
The geometry has a very high aspect ratio. Adjust the View settings to make it easier to see.
Definitions
View 1
In the Model Builder window, expand the Component 1 (comp1)>Definitions node.
Axis
1
In the Model Builder window, expand the View 1 node, then click Axis.
2
In the Settings window for Axis, locate the Axis section.
3
From the View scale list, choose Automatic.
4
Click  Update.
5
Click the  Zoom Extents button in the Graphics toolbar.
Materials
Material 1 (mat1)
1
In the Model Builder window, under Component 1 (comp1) right-click Materials and choose Blank Material.
2
In the Settings window for Material, locate the Material Contents section.
3
Turbulent Flow, k-ε (spf)
Inlet 1
1
In the Model Builder window, under Component 1 (comp1) right-click Turbulent Flow, k-ε (spf) and choose Inlet.
2
In the Settings window for Inlet, locate the Boundary Condition section.
3
From the list, choose Fully developed flow.
4
5
Locate the Fully Developed Flow section. In the Uav text field, type va.
Outlet 1
1
In the Physics toolbar, click  Boundaries and choose Outlet.
2
Symmetry 1
1
In the Physics toolbar, click  Boundaries and choose Symmetry.
2
Mesh 1
Mapped 1
In the Mesh toolbar, click  Mapped.
Distribution 1
1
Right-click Mapped 1 and choose Distribution.
2
3
In the Settings window for Distribution, locate the Distribution section.
4
In the Number of elements text field, type nelemHeight.
Distribution 2
1
In the Model Builder window, right-click Mapped 1 and choose Distribution.
2
3
In the Settings window for Distribution, locate the Distribution section.
4
From the Distribution type list, choose Predefined.
5
In the Number of elements text field, type nelemWidth.
6
In the Element ratio text field, type 15.
7
Click  Build All.
Study 1
In the Home toolbar, click  Compute.
Results
Velocity (spf)
Add some streamlines to the default plot of the fluid velocity.
Streamline 1
1
Right-click Velocity (spf) and choose Streamline.
2
3
In the Settings window for Streamline, locate the Coloring and Style section.
4
Find the Point style subsection. From the Color list, choose White.
5
In the Velocity (spf) toolbar, click  Plot.
Compare the resulting plot to Figure 1.
Turbulence Time Scale
1
In the Home toolbar, click  Add Plot Group and choose 2D Plot Group.
2
In the Settings window for 2D Plot Group, type Turbulence Time Scale in the Label text field.
Surface 1
1
Right-click Turbulence Time Scale and choose Surface.
2
In the Settings window for Surface, click Replace Expression in the upper-right corner of the Expression section. From the menu, choose Component 1 (comp1)>Turbulent Flow, k-ε>Turbulence variables>spf.tauT - Turbulence time scale - s.
3
In the Turbulence Time Scale toolbar, click  Plot.
Compare the resulting plot to Figure 2.
Next, solve for the particle trajectories in the turbulent flow.
Add Physics
1
In the Home toolbar, click  Add Physics to open the Add Physics window.
2
Go to the Add Physics window.
3
In the tree, select Fluid Flow>Particle Tracing>Particle Tracing for Fluid Flow (fpt).
4
At the bottom of the Add Physics section, clear the check box next to Study 1. The particle trajectories are not solved for in the Stationary study step.
5
Click Add to Component 1 in the window toolbar.
6
In the Home toolbar, click  Add Physics to close the Add Physics window.
Add Study
1
In the Home toolbar, click  Add Study to open the Add Study window.
2
Go to the Add Study window.
3
At the bottom of the Add Study section, clear the check box next to the Turbulent Flow, k-ε interface, which will not be solved for in the Time Dependent study step.
4
Find the Studies subsection. In the Select Study tree, select General Studies>Time Dependent.
5
Click Add Study in the window toolbar.
6
In the Home toolbar, click  Add Study to close the Add Study window.
Particle Tracing for Fluid Flow (fpt)
Particle Properties 1
1
In the Model Builder window, under Component 1 (comp1)>Particle Tracing for Fluid Flow (fpt) click Particle Properties 1.
2
In the Settings window for Particle Properties, locate the Particle Properties section.
3
From the ρp list, choose User defined. In the associated text field, type rhop.
4
In the dp text field, type dp.
Assign the Outlet condition to both the top and bottom boundaries. Assign the Symmetry condition at the symmetry axis. The distance from the remaining Wall boundary will be used to compute the anisotropic turbulent velocity perturbations.
Outlet 1
1
In the Physics toolbar, click  Boundaries and choose Outlet.
2
Symmetry 1
1
In the Physics toolbar, click  Boundaries and choose Symmetry.
2
Wall 1
1
In the Model Builder window, click Wall 1.
2
In the Settings window for Wall, locate the Wall Condition section.
3
From the Wall condition list, choose Bounce.
Add the drag force, using the fluid velocity and turbulence variables computed in the previous study.
Drag Force 1
1
In the Physics toolbar, click  Domains and choose Drag Force.
2
3
In the Settings window for Drag Force, locate the Drag Force section.
4
From the u list, choose Velocity field (spf).
5
Locate the Turbulent Dispersion section. From the Turbulent dispersion model list, choose Continuous random walk.
6
From the k list, choose Turbulent kinetic energy (spf).
7
From the ε list, choose Turbulent dissipation rate (spf).
8
Select the Include anisotropic turbulence in boundary layers check box.
9
In the u text field, type ustar_exp.
Release particles from the middle of the channel. Initially, the number density of particles is uniform over the width of the channel.
Release from Grid 1
1
In the Physics toolbar, click  Global and choose Release from Grid.
2
In the Settings window for Release from Grid, locate the Initial Coordinates section.
3
Click  X Range.
4
In the Range dialog box, choose Number of values from the Entry method list.
5
In the Start text field, type halfWidth/(2*Np).
6
In the Stop text field, type halfWidth*(1-1/(2*Np)).
7
In the Number of values text field, type Np.
8
Click Replace.
9
In the Settings window for Release from Grid, locate the Initial Coordinates section.
10
In the qy,0 text field, type height/2.
11
Locate the Initial Velocity section. Specify the v0 vector as
Study 2
Parametric Sweep
1
In the Study toolbar, click  Parametric Sweep.
2
In the Settings window for Parametric Sweep, locate the Study Settings section.
3
4
Step 1: Time Dependent
1
In the Model Builder window, click Step 1: Time Dependent.
2
In the Settings window for Time Dependent, locate the Study Settings section.
3
In the Output times text field, type 0 t1 t2.
4
Click to expand the Values of Dependent Variables section. Find the Values of variables not solved for subsection. From the Settings list, choose User controlled.
5
From the Method list, choose Solution.
6
From the Study list, choose Study 1, Stationary.
7
In the Study toolbar, click  Compute.
Results
Particle Trajectories (fpt)
Compare the default trajectory plot to Figure 3.
Number Density, St = 0.2
1
In the Home toolbar, click  Add Plot Group and choose 1D Plot Group.
2
In the Settings window for 1D Plot Group, locate the Data section.
3
From the Dataset list, choose Particle 1.
4
In the Label text field, type Number Density, St = 0.2.
5
Locate the Data section. From the Parameter selection (St) list, choose From list.
6
In the Parameter values (St) list, select 0.2.
7
From the Time selection list, choose From list.
8
In the Times (s) list, choose 0.76433 and 1.2739.
9
Click to expand the Title section. From the Title type list, choose Manual.
10
In the Title text area, type St = 0.2.
11
Locate the Plot Settings section. Select the x-axis label check box.
12
Histogram 1
1
Right-click Number Density, St = 0.2 and choose Histogram.
2
In the Settings window for Histogram, locate the Expression section.
3
In the Expression text field, type fpt.df1.yplus.
4
Locate the Bins section. In the Number text field, type 50.
5
Click to expand the Legends section. Select the Show legends check box.
6
From the Legends list, choose Manual.
7
8
In the Number Density, St = 0.2 toolbar, click  Plot.
Number Density, St = 0.2
The histogram shows that the number density throughout the cross section is nearly uniform. Use a logarithmic scale and manual axis limits to more easily compare such histograms for each value of the Stokes number.
1
Click the  x-Axis Log Scale button in the Graphics toolbar.
2
Click the  y-Axis Log Scale button in the Graphics toolbar.
3
In the Model Builder window, click Number Density, St = 0.2.
4
In the Settings window for 1D Plot Group, locate the Axis section.
5
Select the Manual axis limits check box.
6
In the x minimum text field, type 1.4.
7
In the x maximum text field, type 150.
8
In the y minimum text field, type 50.
9
In the y maximum text field, type 1e4.
10
In the Number Density, St = 0.2 toolbar, click  Plot.
Duplicate the 1D Plot Group containing the Histogram plot and select other values of the Stokes number to observe how particle inertia affects the particle number density in the channel cross section. As the Stokes number increases, the particles begin to accumulate in the boundary layer close to the wall. All six sets of histograms are shown in Figure 4.