Applied Computational Fluid Dynamics by Hyoung Woo Oh - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

It is presumed that the deviations between observed flow pattern and predicted flow behaviour are caused by the turbulence model used. Some authors already compared different

models that simplify turbulent effects by correlation factors and additional equations like the k-ε RNG model. They stated that the k-ε RNG model is a reasonable compromise between

computational demand and validity of the results (Wardle et al., 2006). Other authors relied on the more extensive Reynolds Stress Model (RSM) or the large-eddy simulation (LES) (Mainza

et al., 2006; Wang & Yu, 2006). Nowakowski et al. reviewed the milestones in the application of computational fluid dynamics in hydrocyclones (Nowakowski et al., 2004). The most common

models used are the k-ε and recently the RSM and LES model. For the evaluation of the

differences in the prediction of the flow behaviour in centrifuges between several turbulence models, the flow pattern was calculated for various rotational speeds and throughputs, using the k-e RNG, k-w, Reynolds Stress Model and large-eddy simulation. A representative

comparison of the obtained results by using the different models is shown in Figure 10. The

plots illustrate the differences in the calculated tangential and axial velocity halfway between inlet and outlet at the angular position of the inlet.

Fig. 10. Tangential (left) and axial velocity profiles (right), throughput 1 l/min, k-e RNG, k-w, RSM and LES turbulence model

index-124_1.jpg

114

Applied Computational Fluid Dynamics

The comparison reveals no significant deviation of the predicted flow pattern. The differences in the tangential velocity are barely detectable. All calculations are in good agreement with the rigid body rotation and thus with the experimentally determined profiles (Spelter, et al. 2011).

Using the LES model, the obtained axial flow profile differs from the other models by

predicting an earlier shift of the main axial flow towards the radius of the outlet at 36 mm.

Because of the high computational demand and instability of the RSM, the simulation with this turbulence model was started with a solution obtained with the k-e RNG model. Nevertheless

the differences are small and all calculated axial flow pattern are in disagreement with the measured flow profile. The mixing and expanding of the inlet jet is significantly understated.

By not questioning and validating the simulations, the user of the CFD software is led to false conclusions which may result in low performance of the centrifuge due to disadvantageous

design. The classical evaluation of the simulation does not indicate a computational error. The mass balance between the inlet and the outlet is fulfilled as stated in Table 1, the residuals are sufficiently small (10-3 to 10-5) and the flow patterns do not change with increasing flow time.

Furthermore the predictions are reproducible with different turbulence models. All these facts suggest an accurate calculation. Only the comparison of the main velocities with

experimentally determined data shows that the prediction misrepresents the true case in

respect to the axial flow profile.

Model

k-e RNG

k-w

RSM LES

% dV

0.0395 0.0015 0.0022 0.0172

Table 1. Deviation of mass balance dV for different turbulence models in % at 3000 rpm and

1 l/min

4.3 Comparison of two and three dimensional modelling

The analysis of the acceleration geometry is only possible in three dimensional modelling,

because, as mentioned in the introduction, the inlet geometry cannot be correctly

reproduced in the two dimensional case. The results presented in chapter 4.1 and 4.2 show a

short-circuit flow between inlet and outlet. The comparison between the results obtained

when calculating three and two dimensional are shown in Figure 11. The predicted

tangential velocity profiles are in good agreement with the rigid body rotation for both cases Fig. 11. Tangential and axial velocity profiles, throughput 1 l/min, k-e RNG, k-w, RSM and

LES turbulence model

index-125_1.jpg

Computational Fluid Dynamics (CFD) and

Discrete Element Method (DEM) Applied to Centrifuges

115

at various rotational speeds. The water is fed through an annulus in the two dimensional

model and the tangential acceleration takes place via the friction at the wall at both sides of the annulus. According to the prediction, this feeding geometry is sufficient for an

acceleration of the incoming liquid up to 9000 rpm. The differences between the cases are

distinguishable in the axial flow profiles which are shown in the diagram on the right side of Figure 11. The profiles presented are the values halfway between inlet and outlet. Due to the symmetry condition, the flow is homogeneously distributed along the whole circumference

for a specific radius. Therefore the maximum axial velocity is lower than in the three

dimensional case due to the conservation of mass. There are recirculating currents next to

the inlet in the two dimensional cases which increase in magnitude with rotational speed. In the three dimensional case, recirculating flows are located next to the inlet as well, but

displaced few degrees in the angle of rotation.

Nevertheless the predicted axial velocity profiles in the two dimensional cases are

comparable to the ones obtained in the three dimensional ones in respect to the radial

distribution of the flow. The inlet jet remains until the water leaves the centrifuge via the circular weir, so that no plug-shaped profile - resulting in a high active volume -, as

observed in the experiments, is formed.

4.4 Calculation of the flow pattern (k-e RNG model, VOF method, transient solution) -

overflow setup

The overflow setup was simulated by using the previously described Volume of Fluid

method. The centrifuge is filled with water at the beginning of the simulation. As an initial condition, the water is already spinning with the rigid body rotation. This assumption is

justifiable because in the experiments, the centrifuge was slowly filled with water and ran

for at least 30 minutes in steady state prior to the measurements. The initial condition

reduces the necessary time until the steady state is reached significantly. Figure 12 shows

Fig. 12. Profiles of water-air interface for a throughput of 1 l/min: a) 1000 rpm; b) 9000 rpm the distribution of the two phases for 1000 (a) und 9000 rpm (b) for a throughput of 1 l/min.

Due to the gravitational and centrifugal force, a conical shape of the water surface is

developed at 1000 rpm. The centrifugal number outbalances the gravitational force with

index-126_1.jpg

index-126_2.jpg

index-126_3.jpg

index-126_4.jpg

index-126_5.jpg

index-126_6.jpg

index-126_7.jpg

index-126_8.jpg

index-126_9.jpg

index-126_10.jpg

index-126_11.jpg

index-126_12.jpg

116

Applied Computational Fluid Dynamics

increasing rotational speed, thus the water phase approaches the shape of a cylinder, as

shown in Figure 12 b).

The tangential velocity profiles for 1000 and 9000 rpm are shown in Figure 13. The diagram

on the right side shows the tangential velocity values averaged over the length of the

centrifuge for the water phase only. The computed values are compared to the rigid body

rotation. The contours of the tangential velocity at 9000 rpm are shown on the left side of

Figure 13. The calculated velocities are in good agreement with the analytical values and

with the measurements. At 9000 rpm, the tangential velocity of the water exhibits a small lag when compared to the rigid body rotation. The magnitude of the deviation is within the

accuracy of the measurements, so that no statement whether this difference is the true case

or not, is possible.

Fig. 13. Tangential velocity profiles, water phase, throughput of 1 l/min; left: contours at 9000 rpm; right: for different rotational speeds and different axial positions

Figure 14 shows the axial velocity profiles for 9000 rpm at a constant throughput of 1 l/min.

The axial flow pattern is similar to the one predicted in the single-phase model. The inlet jet Fig. 14. Axial velocity profiles at 9000 rpm and 1 l/min throughput; left: contours; right:

different distances from the inlet

index-127_1.jpg

Computational Fluid Dynamics (CFD) and

Discrete Element Method (DEM) Applied to Centrifuges

117

does not widen so that a short-circuit flow towards the outlet is formed, as shown in the

contour plot on the left side of Figure 14. There is no significant expansion of the flow

profile neither in angular nor in radial direction. The peak of the axial velocity shifts from a radius of 42 mm close to the inlet towards a radius of 36 mm. This behaviour was also

predicted in the single-phase model. Nevertheless the simulation misrepresents the true

case, in which a significant widening of the inlet jet was observed.

4.5 Influence of the inlet geometry on the flow pattern

In order to investigate how the acceleration of the feed affects the flow pattern, a centrifuge with a simpler geometry and a low performance feed accelerator was simulated. The main

differences when compared to the tubular bowl centrifuge are the lower length-to-diameter

ratio of 0.8 and the feed accelerator. The feed enters a rotating disc and leaves it radially with a certain angular velocity. The water exits the centrifuge through eight boreholes

distributed along the circumference at the top of the bowl. Figure 15 shows a scheme of the

centrifuge geometry on the left and the results of the volume fraction of air on the right.

Water is fed axially through the inlet to the accelerator; there, the water changes its direction and gains in tangential velocity. The transfer of rotational movement occurs just at the plate surfaces, defined as no-slip boundaries, via momentum diffusion. Then, the water exits the

accelerator as streams that travel through the air core and enter the rotating liquid pool. The radius of the interface changes 35 % along the height of the centrifuge from the inlet to the upper part of the centrifuge because of the low rotational speed (454 g).

Fig. 15. Left: geometry of the centrifuge; right: contours of volume fraction of air at 2550 rpm and 5.8 l/min

The main flow occurs in the direction of the rotation of the bowl. Both, the liquid pool and the air core rotate. The tangential velocity of the water jet from the inlet accelerator (height 0.0275 m) is lower than the tangential velocity of the bowl. This causes a drag of the whole rotating liquid pool regarding the angular velocity of the bowl. The values of the tangential velocity of the liquid averaged over the circumference are plotted for different heights in

Figure 16. The averaged velocity of the liquid in this geometry is approximately 50 % below

the rigid body rotation. The values at the wall have to match the rigid body rotation of the bowl to fulfil the no-slip boundary condition. The steep decrease of the tangential velocity

index-128_1.png

118

Applied Computational Fluid Dynamics

near the wall was found to be smoother for a 6 times finer grid. However, the computational

effort for the multiphase simulation including the particles with the finer grid was

unaffordable. The deviation with respect to the rigid body rotation decreases for lower flow rates because a smaller amount of liquid has to be accelerated. The simulation of this

geometry shows how important is to accelerate the feed properly to the angular velocity of

the liquid pool before entering it in order to reach the angular velocity of the bowl in the rotating liquid and thus the highest centrifugal acceleration possible.

In contrast to the other centrifuge, a layer with high axial velocity at the gas-liquid interface, as expected by the boundary layer model, has been observed in the simulations as depicted

in the diagram on the right in Figure 16. Near the impact height of the inlet water jet two

whirls with opposed directions are formed. This can be seen at the negative axial velocity

values at the interface for the height 0.02 m, just below the inlet, and the positive axial

velocities at he interface for the height 0.0275 m, just above the inlet. Upsides the height of the inlet, a homogeneous axial boundary layer is formed at the interface with a

recirculation layer at the bowl wall. This layer exhibits a similar averaged thickness δ as

the one proposed by the theoretical model of Gosele (Gösele, 1968) but much thicker as

the one calculated following the theory of Reuter (Reuter, 1967) as seen in Table 2. Indeed, it occupies the whole liquid pool until a recirculation layer is formed close to the bowl

wall. This is due to the lower tangential velocities which cause a small pressure gradient

in the radial direction reducing the stratification of the flow and allowing the fluid to

move in the axial direction. The velocity values oscillate with the angle, especially near

the inlet at 0.02 m and 0.0275 m and the outlet at 0.0975 m, respectively. Thus the standard deviation is higher at these positions.

Fig. 16. Left: tangential velocity versus radial position; right: axial velocity versus radial position at 2550 rpm and 5.8 l/min

δsimulated (mm)

δReuter (mm)

δGösele (mm)

14.1 1.71

15.8

Table 2. Thickness of the axial boundary layer δ obtained in the simulations and calculated

analytically at 2550 rpm and a throughput of 5.8 l/min

Computational Fluid Dynamics (CFD) and

Discrete Element Method (DEM) Applied to Centrifuges

119

5. Mathematical formulation of the particles

There are several multiphase models available for simulation of a particulate phase. A

summary of the different simulation methods for multiphase flows in CFD can be found in

(van Wachem & Almstedt, 2003). The multiphase models can be divided into two groups;

the Euler-Euler and the Euler-Lagrange methods.

The Euler-Euler method considers all phases as continuous ones. For each phase, a

momentum conservation equation is solved. In the momentum equation of the dispersed

phases, the kinetic theory for granular flows (Lun & Savage, 1987) is used. The intensity of the particle velocity fluctuations determines the stresses, viscosity, and pressure of the solid phase. The interaction between phases is considered by means of momentum exchange

coefficients. This numerical method has the advantage that it is suitable for high volume

concentrations of the disperse phase. The disadvantage is that the properties of each

dispersed particle cannot be simulated and a particle size distribution cannot be considered.

Moreover, the stability of the calculation depends on the choice of model parameters and

often convergence problems appear. For the simulation of sedimentation and thickening

processes there are also some Euler-Euler models based on the Kynch’s kinematics

sedimentation theory (Kynch, 1952), which were reviewed by Bürger in (Bürger &

Wendland, 2001). These models are all based on the measurement of fundamental material

properties of the suspension to be separated as explained by (Buscall, 1990; Buscall et al., 1987; Landman & White, 1994). These multiphase models are based on mass balances for the fluid and the disperse phase. The fluid flow is not calculated but they consider the relative velocity between fluid and particles. On this basis, these models are unsuitable for studying the particle behaviour in solid bowl centrifuges where the liquid flow plays a crucial role in particle settling. These models usually describe the sedimentation and consolidation in

batch processes in simple-geometry centrifuges. An extension of these models is necessary

to describe the sediment build-up in two and three dimensions correctly.

The Euler-Lagrange method solves the continuum conservation equations just for the

characterization of the continuous phase. It describes the dispersed phase as mass points

and determines the particle trajectories by integrating a force balance for each individual

particle. This way, instead of a partial differential equation, only an ordinary differential equation must be solved for the dispersed phase. The weakness of the Euler-Lagrange

method is that it is only valid for diluted suspensions if a one-way or a two-way coupling

method between continuous and disperse phase is applied. With the one-way coupling, the

motion of the discrete phase is calculated based on the velocity field of the continuous phase but the particles have no effect on the pressure and velocity field of the continuous phase. By using a two-way coupling, the equations of continuous and discrete phase are calculated

alternatively. The momentum conservation equation of the continuous phase has a source

term to consider the momentum exchange with the particulate phase. The maximum value for

the volume concentration of the disperse phase is between 5% (Schütz et al., 2007) and 10-12%

(ANSYS, 2009), above this value the interaction between particles should be considered. Thus there is a third method, four-way coupling, which involves the interaction within the disperse phase: impacts between particles, particle-wall interactions, van der Waals forces, electrostatic forces, etc. Thus the restriction for the volume fraction is not valid anymore. However, if the number of particles increases, the computational effort rises significantly. Usually the

characteristics of the problem and the information that should be obtained from the simulation decide which multiphase model and coupling methodology should be used.

index-130_1.jpg

index-130_2.jpg

index-130_3.jpg

index-130_4.jpg

120

Applied Computational Fluid Dynamics

The Discrete Element Method (DEM) (Cundall & Strack, 1979) is a Lagrangian method

originally developed for the simulation of bulk solids and then extended to the entire field of particle technology. This method solves the Newton's equations of motion for translation

and rotation of the particles. In these equations, impact forces of particle-particle and

particle-wall interactions are considered. Other forces such as electrostatic or Brownian

forces can also be implemented if necessary. In order to describe the collisions of particles, the soft-sphere approach was chosen. When two particles collide, they actually deform. This

deformation is described by an overlap displacement of the particles. A schematic

representation of the forces on the particles in the soft-sphere model is depicted in Figure 17.

This approach uses a spring-dashpot-slider system to calculate the behaviour of the particles during the contact time. The soft-sphere model allows contacts of a particle with more

neighbouring particles during a time step. The net contact force acting on a particle is added in the case of multiple overlapping. This model for contact forces was first developed by

Cundall (Cundall & Strack, 1979), who proposed a parallel linear spring-dashpot model for the normal direction and a parallel linear spring-dashpot in a series with a slider for the

tangential direction. A comparison of the performance of several soft-sphere models can be

found in (Di Renzo & Di Maio, 2005; Stevens & Hrenya, 2005). There, also the Hertz-Mindlin (Mindlin, 1949) non-linear soft-sphere model used in the present work is analysed. This

model proposes a normal force as a function of the normal overlap δn, the Young’s modulus

of the particle material and the particle radii. The damping force in normal direction

depends on the restitution coefficient, the normal stiffness, and the normal component of the relative velocity between the particles. The tangential force depends on the tangential

overlap δt, on the shear modulus of the particle material, and on the particle radii; while in the damping force in tangential direction the tangential stiffness and the tangential

component of the relative velocity between the particles are considered.

Fig. 17. Contact forces between two particles following the soft-sphere model

While the DEM model was originally used in environments without fluid, increasing efforts

have been made to extend this model to the mixture of fluid and particles. For that, a coupling of the DEM calculations with the fluid flow simulation is necessary to be able to

include the hydrodynamic forces in the particle simulation. An important aspect in the

calculation of particles in a fluid is the energy dissipation into the surrounding fluid. In viscous fluids, there are, in contrast to air, two additional effects that lead to energy losses.

The drag force leads to lower particle velocities and when a particle is about to hit a surface or another particle, a deceleration occurs. This is due to the drainage of the fluid, which is located in the gap between the two particles or the particle and the surface. In his work

index-131_1.png

index-131_2.png

index-131_3.png

Computational Fluid Dynamics (CFD) and

Discrete Element Method (DEM) Applied to Centrifuges

121

(Gondret et al., 2002), Gondret presents the coefficient of restitution as a function of the Stokes number St (Equation 16), which describes the ratio of the dynamic response time of a

particle to flow changes and the characteristic flow time. He determined empirically a

critical Stokes number of about 10. Below this value, a particle does not bounce when it hits a wall and instead remains in contact with it.

1 sv x

St 

(16)

9

In a centrifugal field, the hydrodynamic forces (drag, lift, torque) play a crucial role and must be included in the calculation of the particle trajectories. Therefore the fluid flow and particles movement have to be computed alternatively. A scheme of the coupling between

the CFD software FLUENT and the DEM software EDEM can be seen in Figure 18. At first,

the flow will be determined with CFD, then for each position in the computational domain,

the velocity, pressure, density and viscosity of the flow are transferred to DEM in order to calculate the hydrodynamic forces and include them in the balance. Afterwards, the fluid

flow will be computed again considering the influence of the particles as an additional sink term in the momentum equations. This sink term include the force transmitted to the

particles by the fluid.

Fig. 18. Coupling between CFD and DEM

Some authors have already simulated such complex multiphase flows in other separation

equipment such as filters and hydrocyclones (Chu et al., 2009; Chu & Yu, 2008; Dong et al., 2003; Ni et al., 2006). Nevertheless due to the high velocity gradients and complex

interactions between the phases, the simulation of multiphase flow and in particular the

sediment behaviour in centrifugal field is still an unsolved challenge. The advantage of this approach is that an accurate description of the flow, the particle paths and the sediment

behaviour can be obtained. However, the amount of simulated particles and its size is

limited due to the computational demand.

5.1 Particles simulation parameters

The calculation in CFD must be run as an unsteady case when coupled with DEM. The DEM

time steps are typically smaller than the CFD time steps in order to correctly capture the

contact behaviour. The choice of time steps in DEM simulation is of great importance in

order to achieve stability in the calculation but does not perform very long calculations,

although still representing the real particle behaviour. Therefore it is necessary to analyse 122

Applied Computational Fluid Dynamics

the forces acting on the particles to determine the different time scales. In the case presented, the Hertzian contact time - the total time of contact in Hertz theory of elastic collision -, the particle sedimentation time - the time required for a particle to settle one diameter -, and the particle relaxation time - the time for a particle to adapt to the surrounding fluid flow - are considered. On the basis of these, the smallest time step must be chosen so that all physical phenomena are represented. In the simulations presented, the Hertzian contact time is the

smallest time scale and, thus a time step tDEM = 10-6 s was chosen in the DEM softwa